This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\AtAppendix\AtAppendix

Collisions of the supercritical Keller-Segel particle system

Nicolas Fournier and Yoan Tardy Sorbonne Université, LPSM-UMR 8001, Case courrier 158,75252 Paris Cedex 05, France. [email protected], [email protected]
Abstract.

We study a particle system naturally associated to the 22-dimensional Keller-Segel equation. It consists of NN Brownian particles in the plane, interacting through a binary attraction in θ/(Nr)\theta/(Nr), where rr stands for the distance between two particles. When the intensity θ\theta of this attraction is greater than 22, this particle system explodes in finite time. We assume that N>3θN>3\theta and study in details what happens near explosion. There are two slightly different scenarios, depending on the values of NN and θ\theta, here is one: at explosion, a cluster consisting of precisely k0k_{0} particles emerges, for some deterministic k07k_{0}\geq 7 depending on NN and θ\theta. Just before explosion, there are infinitely many (k01)(k_{0}-1)-ary collisions. There are also infinitely many (k02)(k_{0}-2)-ary collisions before each (k01)(k_{0}-1)-ary collision. And there are infinitely many binary collisions before each (k02)(k_{0}-2)-ary collision. Finally, collisions of subsets of 3,,k033,\dots,k_{0}-3 particles never occur. The other scenario is similar except that there are no (k02)(k_{0}-2)-ary collisions.

Key words and phrases:
Keller-Segel equation, Stochastic particle systems, Bessel processes, Collisions
2010 Mathematics Subject Classification:
60H10, 60K35
The fruitful comments of the two referees helped us to substantially improve the presentation of the paper.

1. Introduction and main results

1.1. Informal definition of the model

We consider some scalar parameter θ>0\theta>0 and a number N2N\geq 2 of particles with positions Xt=(Xt1,,XtN)(2)NX_{t}=(X^{1}_{t},\dots,X^{N}_{t})\in({\mathbb{R}}^{2})^{N} at time t0t\geq 0. Informally, we assume that the dynamics of these particles are given by the system of S.D.E.s

(1) dXti=dBtiθNjiXtiXtjXtiXtj2dt,i[[1,N]],\displaystyle{\rm d}X^{i}_{t}={\rm d}B^{i}_{t}-\frac{\theta}{N}\sum_{j\neq i}\frac{X^{i}_{t}-X^{j}_{t}}{\|X^{i}_{t}-X^{j}_{t}\|^{2}}{\rm d}t,\qquad i\in[\![1,N]\!],

where the 22-dimensional Brownian motions ((Bti)t0)i[[1,N]]((B^{i}_{t})_{t\geq 0})_{i\in[\![1,N]\!]} are independent. In other words, we have NN Brownian particles in the plane interacting through an attraction in 1/r1/r, which is Coulombian in dimension 22. Actually, this S.D.E. does not clearly make sense, due to the singularity of the drift, and we will use, as suggested by Cattiaux-Pédèches [4], the theory of Dirichlet spaces, see Fukushima-Oshima-Takeda [11].

1.2. Brief motivation and informal presentation of the main results

This particle system is very natural from a physical point of view, because, as we will see, there is a tight competition between the Brownian excitation and the Coulombian attraction. It can also be seen as an approximation of the famous Keller-Segel equation [16], see also Patlak [20]. This nonlinear P.D.E. has been introduced to model the collective motion of cells, which are attracted by a chemical substance that they emit. It is well-known that a phase transition occurs: if the intensity of the attraction is small, then there exist global solutions, while if the attraction is large, the solution explodes in finite time.

We will show that this phase transition already occurs at the level of the particle system (1): there exist global (very weak) solutions if θ(0,2)\theta\in(0,2) (subcritical case, see Proposition 1.3 below), but solutions must explode in finite time if θ2\theta\geq 2 (supercritical case).

To our knowledge, the supercritical case has not been studied in details, and we aim to describe precisely the explosion phenomenon. Informally, we will show the following (see Theorem 1.5 below). We assume that θ2\theta\geq 2 and N>3θN>3\theta, we set k0=2N/θ[[7,N]]k_{0}=\lceil 2N/\theta\rceil\in[\![7,N]\!]. There exists a (very weak) solution (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta)} to (1), with ζ<\zeta<\infty a.s. and such that Xζ=limtζXtX_{\zeta-}=\lim_{t\to\zeta-}X_{t} exists. Moreover, there is a cluster containing precisely k0k_{0} particles in the configuration XζX_{\zeta-}, and no cluster containing strictly more than k0k_{0} particles. Such a cluster containing k0k_{0} particles is inseparable, so that (1) is meaningless (even in a very weak sense) after ζ\zeta. Just before explosion, there are infinitely many k1k_{1}-ary collisions, where k1=k01k_{1}=k_{0}-1. If (k03)(2(k02)θ/N)<2(k_{0}-3)(2-(k_{0}-2)\theta/N)<2, we set k2=k12k_{2}=k_{1}-2 and just before each k1k_{1}-ary collision, there are infinitely many k2k_{2}-collisions. Else, we set k2=k1k_{2}=k_{1}. In any case, there are infinitely many binary collisions just before each k2k_{2}-ary collision. During the whole time interval [0,ζ)[0,\zeta), there are no kk-ary collisions, for any k[[3,k21]]k\in[\![3,k_{2}-1]\!].

This phenomenon seems surprising and original, in particular because of the gap between binary and k2k_{2}-ary collisions.

1.3. Sets of configurations

We introduce, for all K[[1,N]]K\subset[\![1,N]\!] and all x=(x1,,xN)(2)Nx=(x^{1},\dots,x^{N})\in({\mathbb{R}}^{2})^{N},

SK(x)=1|K|iKxi2andRK(x)=iKxiSK(x)2=12|K|i,jKxixj20.\displaystyle S_{K}(x)=\frac{1}{|K|}\sum_{i\in K}x^{i}\in{\mathbb{R}}^{2}\quad\hbox{and}\quad R_{K}(x)=\sum_{i\in K}\|x^{i}-S_{K}(x)\|^{2}=\frac{1}{2|K|}\sum_{i,j\in K}\|x^{i}-x^{j}\|^{2}\geq 0.

Here |K||K| is the cardinal of KK and \|\cdot\| stands for the Euclidean norm in 2{\mathbb{R}}^{2}. Observe that RK(x)=0R_{K}(x)=0 if and only if all the particles indexed in KK are at the same place. We also set, for k2k\geq 2,

Ek={x(2)N:K[[1,N]] with cardinal |K|=k,RK(x)>0},E_{k}=\Big{\{}x\in({\mathbb{R}}^{2})^{N}:\forall K\subset[\![1,N]\!]\mbox{ with cardinal }|K|=k,\;R_{K}(x)>0\Big{\}},

which represents the set of configurations with no cluster of kk (or more) particles. Observe that Ek=(2)NE_{k}=({\mathbb{R}}^{2})^{N} for all k>Nk>N.

1.4. Bessel processes

We recall that a squared Bessel process (Zt)t0(Z_{t})_{t\geq 0} of dimension δ\delta\in{\mathbb{R}} is a nonnegative solution, killed when it reaches 0 if δ0\delta\leq 0, of the equation

Zt=Z0+20tZsdWs+δt,Z_{t}=Z_{0}+2\int_{0}^{t}\sqrt{Z_{s}}{\rm d}W_{s}+\delta t,

where (Wt)t0(W_{t})_{t\geq 0} is a 11-dimensional Brownian motion. We then say that (Zt)t0(\sqrt{Z_{t}})_{t\geq 0} is a Bessel process of dimension δ\delta. This process has the following property, see Revuz-Yor [21, Chapter XI]: \bullet if δ2\delta\geq 2, then a.s., for all t>0t>0, Zt>0Z_{t}>0;

\bullet if δ(0,2)\delta\in(0,2), then a.s., ZZ is reflected infinitely often at 0; \bullet if δ0\delta\leq 0, then ZZ a.s. hits 0 and is then killed.

Applying informally the Itô formula, one finds that Yt=ZtY_{t}=\sqrt{Z_{t}} should solve

Yt=Y0+Wt+δ120tdsYs,Y_{t}=Y_{0}+W_{t}+\frac{\delta-1}{2}\int_{0}^{t}\frac{{\rm d}s}{Y_{s}},

which resembles (1) in that we have a Brownian excitation in competition with an attraction by 0, or a repulsion by 0, depending on the value of δ\delta, proportional to 1/r1/r. This formula rigorously holds true only when δ>1\delta>1, see [21, Chapter XI].

1.5. Some important quantities

Consider a (possibly very weak) solution (Xt)t0(X_{t})_{t\geq 0} to (1). As we will see, when fixing a subset K[[1,N]]K\subset[\![1,N]\!] and when neglecting the interactions between the particles indexed in KK and the other ones, one finds that the process (RK(Xt))t0(R_{K}(X_{t}))_{t\geq 0} behaves like a squared Bessel process with dimension dθ,N(|K|)d_{\theta,N}(|K|), where

(2) dθ,N(k)=(k1)(2kθN).d_{\theta,N}(k)=(k-1)\Big{(}2-\frac{k\theta}{N}\Big{)}.

Similar computations already appear in Haškovec-Schmeiser [12], see also [8]. A little study, see Appendix A, see also Figure 1 and Subsection 1.8 for numerical examples, shows the following facts. For r+r\in{\mathbb{R}}_{+}, we set r=min{n:nr}\lceil r\rceil=\min\{n\in{\mathbb{N}}:n\geq r\}.

Lemma 1.1.

Fix θ>0\theta>0 and N2N\geq 2 such that N>θN>\theta. For k0=2Nθ3k_{0}=\lceil\frac{2N}{\theta}\rceil\geq 3, we have

(3) dθ,N(k)>0ifk[[2,k01]]anddθ,N(k)0ifkk0.d_{\theta,N}(k)>0\quad\hbox{if}\quad k\in[\![2,k_{0}-1]\!]\quad\hbox{and}\quad d_{\theta,N}(k)\leq 0\quad\hbox{if}\quad k\geq k_{0}.

We also define k1=k01k_{1}=k_{0}-1, and

k2={k02 if dθ,N(k02)<2,k01 if dθ,N(k02)2.k_{2}=\left\{\begin{array}[]{lll}k_{0}-2&\hbox{ if }&d_{\theta,N}(k_{0}-2)<2,\\[4.0pt] k_{0}-1&\hbox{ if }&d_{\theta,N}(k_{0}-2)\geq 2.\end{array}\right.

If θ2\theta\geq 2 and N>3θN>3\theta, then k0[[7,N]]k_{0}\in[\![7,N]\!] and it holds that \bullet dθ,N(2)(0,2)d_{\theta,N}(2)\in(0,2); \bullet dθ,N(k)2d_{\theta,N}(k)\geq 2 if k[[3,k21]]k\in[\![3,k_{2}-1]\!]; \bullet dθ,N(k)(0,2)d_{\theta,N}(k)\in(0,2) if k{k2,k1}k\in\{k_{2},k_{1}\}; \bullet dθ,N(k)0d_{\theta,N}(k)\leq 0 if kk0k\geq k_{0}.

Figure 1. Plot of dθ,N(k)d_{\theta,N}(k) as a function of k[[2,N]]k\in[\![2,N]\!] with N=9N=9 and with θ=2.35\theta=2.35 (left) and θ=2.42\theta=2.42 (right). [Uncaptioned image] k0=8k_{0}=8, k1=7k_{1}=7, k2=7k_{2}=7 k0=8k_{0}=8, k1=7k_{1}=7, k2=6k_{2}=6

We thus expect that there may be some non sticky kk-ary collisions for k{2,k2,k1}k\in\{2,k_{2},k_{1}\}, some sticky kk-ary collisions when kk0k\geq k_{0}, but no kk-ary collision for k[[3,k21]]k\in[\![3,k_{2}-1]\!].

1.6. Generator and invariant measure

As we will see in Subsection 3.13, the S.D.E. (1) cannot have a solution in the classical sense, at least when dθ,N(k1)(0,1)d_{\theta,N}(k_{1})\in(0,1), because the drift term cannot be integrable in time. We will thus define a solution through the theory of the Dirichlet spaces.

For x=(x1,,xN)(2)Nx=(x^{1},\dots,x^{N})\in({\mathbb{R}}^{2})^{N} and for dx{\rm d}x the Lebesgue measure on (2)N({\mathbb{R}}^{2})^{N}, we set

(4) 𝐦(x)=1ijNxixjθ/Nandμ(dx)=𝐦(x)dx,{\mathbf{m}}(x)=\prod_{1\leq i\neq j\leq N}\|x^{i}-x^{j}\|^{-\theta/N}\quad\hbox{and}\quad\mu({\rm d}x)={\mathbf{m}}(x){\rm d}x,

where {1ijN}\{1\leq i\neq j\leq N\} stands for the set {(i,j)[[1,N]]2:ij}\{(i,j)\in[\![1,N]\!]^{2}:i\neq j\}. Informally, the generator of the solution to (1) is given by X{\mathcal{L}}^{X}, where for φC2((2)N)\varphi\in C^{2}(({\mathbb{R}}^{2})^{N}),

(5) Xφ(x)=12Δφ(x)θN1ijNxixjxixj2xiφ(x)=12𝐦(x)div[𝐦(x)φ(x)],{\mathcal{L}}^{X}\varphi(x)=\frac{1}{2}\Delta\varphi(x)-\frac{\theta}{N}\sum_{1\leq i\neq j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}}\cdot\nabla_{x^{i}}\varphi(x)=\frac{1}{2{\mathbf{m}}(x)}{\rm div}[{\mathbf{m}}(x)\nabla\varphi(x)],

see (4) for the last equality. It is well-defined for all xE2x\in E_{2} and μ\mu-symmetric. Indeed, an integration by parts shows that

(6) φ,ψCc2(E2),(2)NφXψdμ=12(2)Nφψdμ=(2)NψXφdμ.\forall\;\varphi,\psi\in C^{2}_{c}(E_{2}),\quad\int_{({\mathbb{R}}^{2})^{N}}\varphi{\mathcal{L}}^{X}\psi\;{\rm d}\mu=-\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi\cdot\nabla\psi\;{\rm d}\mu=\int_{({\mathbb{R}}^{2})^{N}}\psi{\mathcal{L}}^{X}\varphi\;{\rm d}\mu.

As we will see in Proposition A.1, the measure μ\mu is Radon on (2)N({\mathbb{R}}^{2})^{N} in the subcritical case θ(0,2)\theta\in(0,2), while it is Radon on Ek0E_{k_{0}} (and not on Ek0+1E_{k_{0}+1}) in the supercritical case θ2\theta\geq 2. This will allow us to use some results found in Fukushima-Oshima-Takeda [11] and to obtain the following existence result.

Proposition 1.2.

We fix N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta and recall that k0=2N/θk_{0}=\lceil 2N/\theta\rceil. We set 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} and 𝒳=𝒳{}{\mathcal{X}}_{\triangle}={\mathcal{X}}\cup\{\triangle\}, where \triangle is a cemetery point. There exists a diffusion 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳)\mathbb{X}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}_{x}^{X})_{x\in{\mathcal{X}_{\triangle}}}) with values in 𝒳{\mathcal{X}}_{\triangle}, which is μ\mu-symmetric, with regular Dirichlet space (X,X)({\mathcal{E}}^{X},{\mathcal{F}}^{X}) on L2((2)N,μ)L^{2}(({\mathbb{R}}^{2})^{N},\mu) with core Cc(𝒳)C^{\infty}_{c}({\mathcal{X}}) defined by

for allφCc(𝒳),X(φ,φ)=12(2)Nφ2dμ=(2)NφXφdμ\hbox{for all}\quad\varphi\in C^{\infty}_{c}({\mathcal{X}}),\quad\mathcal{E}^{X}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\|\nabla\varphi\|^{2}{\rm d}\mu=-\int_{({\mathbb{R}}^{2})^{N}}\varphi{\mathcal{L}}^{X}\varphi\;{\rm d}\mu

and such that for all xE2x\in E_{2}, all t>0t>0, the law of XtX_{t} under x{\mathbb{P}}_{x} has a density with respect to the Lebesgue measure on (2)N({\mathbb{R}}^{2})^{N}. We call such a process a KS(θ,N)KS(\theta,N)-process and denote by ζ=inf{t0:Xt=}\zeta=\inf\{t\geq 0:X_{t}=\triangle\} its life-time.

We refer to Subsection B.1 for a quick summary about the notions used in this proposition: diffusion (i.e. continuous Hunt process), link between its generator, semi-group and Dirichlet space, definition of the one-point compactification topology endowing 𝒳{\mathcal{X}}_{\triangle}, etc. Let us mention that by definition, \triangle is absorbing, i.e. Xt=X_{t}=\triangle for all tζt\geq\zeta. Also, tXtt\mapsto X_{t} is a priori continuous on [0,)[0,\infty) only for the one-point compactification topology on 𝒳{\mathcal{X}_{\triangle}}, which precisely means that it is continuous for the usual topology of (2)N({\mathbb{R}}^{2})^{N} during [0,ζ)[0,\zeta), and it holds that ζ=limninf{t0:Xt𝒦n}\zeta=\lim_{n\to\infty}\inf\{t\geq 0:X_{t}\notin{\mathcal{K}}_{n}\} for any increasing sequence of compact subsets (𝒦n)n1({\mathcal{K}}_{n})_{n\geq 1} of Ek0E_{k_{0}} such that n1𝒦n=Ek0\cup_{n\geq 1}{\mathcal{K}}_{n}=E_{k_{0}}.

As we will see in Remark 11.6, for all xE2x\in E_{2}, under xX{\mathbb{P}}^{X}_{x}, XtX_{t} solves (1) during [0,σ)[0,\sigma), where σ=inf{t0:XtE2}\sigma=\inf\{t\geq 0:X_{t}\notin E_{2}\}. By the Markov property, this implies XtX_{t} solves (1) during any open time-interval on which it does not visit 𝒳E2{\mathcal{X}}\setminus E_{2}.

When θ<2\theta<2, we have k0>Nk_{0}>N and thus Ek0=(2)NE_{k_{0}}=({\mathbb{R}}^{2})^{N}. We will easily prove the following non-explosion result, which is almost contained in Cattiaux-Pédèches [4], who treat the case where θ(0,2(N2)/(N1))\theta\in(0,2(N-2)/(N-1)).

Proposition 1.3.

Fix θ(0,2)\theta\in(0,2) and N2N\geq 2. Consider the KS(θ,N)KS(\theta,N)-process 𝕏{\mathbb{X}} introduced in Proposition 1.2. For all xE2x\in E_{2}, we have x(ζ=)=1{\mathbb{P}}_{x}(\zeta=\infty)=1.

When θ2\theta\geq 2, we will see that there is explosion. Note that any collision of a set of kk0k\geq k_{0} particles makes the process leave Ek0E_{k_{0}} and thus explode. However, it is not clear at all at this point that explosion is due to a precise collision: the process could explode because it tends to infinity (which is not hard to exclude) or to the boundary of Ek0E_{k_{0}} with possibly many oscillations.

1.7. Main result

To avoid any confusion, let us define precisely what we call a collision.

Definition 1.4.

(i) For K[[1,N]]K\subset[\![1,N]\!], we say that there is a KK-collision in the configuration x(2)Nx\in({\mathbb{R}}^{2})^{N} if RK(x)=0R_{K}(x)=0 and if RK{i}(x)>0R_{K\cup\{i\}}(x)>0 for all i[[1,N]]Ki\in[\![1,N]\!]\setminus K.

(ii) For a (2)N({\mathbb{R}}^{2})^{N}-valued process (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta)}, we say that there is a KK-collision at time s[0,ζ)s\in[0,\zeta) if there is a KK-collision in the configuration XsX_{s}.

The main result of this paper is the following description of the explosion phenomenon.

Theorem 1.5.

Assume that θ2\theta\geq 2, that N>3θN>3\theta and recall that k0[[7,N]]k_{0}\in[\![7,N]\!], k1=k01k_{1}=k_{0}-1 and k2{k01,k02}k_{2}\in\{k_{0}-1,k_{0}-2\} were defined in Lemma 1.1. Consider the KS(θ,N)KS(\theta,N)-process 𝕏{\mathbb{X}} introduced in Proposition 1.2. For all xE2x\in E_{2}, we x{\mathbb{P}}_{x}-a.s. have the following properties: (i) ζ\zeta is finite and Xζ=limtζXtX_{\zeta-}=\lim_{t\rightarrow\zeta-}X_{t} exists for the usual topology of (2)N({\mathbb{R}}^{2})^{N}; (ii) there is K0[[1,N]]K_{0}\subset[\![1,N]\!] with cardinal |K0|=k0|K_{0}|=k_{0} such that there is a K0K_{0}-collision in the configuration XζX_{\zeta-}, and for all K[[1,N]]K\subset[\![1,N]\!] such that |K|>k0|K|>k_{0}, there is no KK-collision in the configuration XζX_{\zeta-}; (iii) for all t[0,ζ)t\in[0,\zeta) and all KK0K\subset K_{0} with cardinal |K|=k1|K|=k_{1}, there is an infinite number of KK-collisions during (t,ζ)(t,\zeta) and none of these instants of KK-collision is isolated; (iv) if k2=k02k_{2}=k_{0}-2, then for all LKK0L\subset K\subset K_{0} such that |L|=k2|L|=k_{2} and |K|=k1|K|=k_{1}, for all instant t(0,ζ)t\in(0,\zeta) of KK-collision and all s[0,t)s\in[0,t), there is an infinite number of LL-collisions during (s,t)(s,t) and none of these instants of LL-collision is isolated; (v) for all K[[1,N]]K\subset[\![1,N]\!] with cardinal |K|[[3,k21]]|K|\in[\![3,k_{2}-1]\!], there is no KK-collision during [0,ζ)[0,\zeta); (vi) for all LKK0L\subset K\subset K_{0} such that |L|=2|L|=2 and |K|=k2|K|=k_{2}, for all instant t(0,ζ)t\in(0,\zeta) of KK-collision and all s[0,t)s\in[0,t), there is an infinite number of LL-collisions during (s,t)(s,t) and none of these instants of LL-collision is isolated.

The condition θ2\theta\geq 2 is crucial to guarantee that k0Nk_{0}\leq N. On the contrary, we impose N>3θN>3\theta for simplicity, because Lemma 1.1 does not hold true without this assumption. The other cases may also be studied, but we believe this is not very restrictive: NN is thought as very large when compared to θ\theta, at least as far as the approximation of the Keller-Segel equation is concerned.

1.8. Comments

Let us mention that the very precise values of NN and θ\theta influence the value k2k_{2}.

(a) If N=200N=200 and θ=4.04\theta=4.04, we have k0=100k_{0}=100, k1=99k_{1}=99 and k2=98k_{2}=98.

(b) If N=200N=200 and θ=4.015\theta=4.015, we have k0=100k_{0}=100 and k1=k2=99k_{1}=k_{2}=99.

Let us describe informally, in the chronological order, what happens e.g. in case (b) above. We start with 200200 particles at 200200 different places. During the whole story, there is no kk-ary collision for k=3,,98k=3,\dots,98. Here and there, two particles meet, they collide an infinite number of times, but manage to separate. Then at some times, we have 9898 particles close to each other and there are many binary collisions. Then, if a 9999-th particle arrives in the same zone (and this eventually occurs), there are infinitely many 9999-ary collisions, with infinitely many binary collisions of all possible pairs before each. These 9999 particles may manage to separate forever, or for a large time, but if a 100100-th particle arrives in the zone (and this situation eventually occurs), then there are infinitely many 9999-ary collisions of all the possible subsets and, finally, a 100100-ary collision producing explosion, and the story is finished. Informally, the resulting cluster is not able to separate, because the attraction dominates the Brownian excitation, since a Bessel process of dimension dθ,N(100)0d_{\theta,N}(100)\leq 0 is absorbed when it reaches 0. We hope to be able, in a future work, to propose and justify a model describing what happens after explosion.

1.9. References

In many papers about the Keller-Segel equation, the parameter χ=4πθ\chi=4\pi\theta is used, so that the transition at θ=2\theta=2 corresponds to the transition at χ=8π\chi=8\pi. As already mentioned, this nonlinear P.D.E. has been introduced to model the collective motion of cells, which are attracted by a chemical substance that they emit. It describes the density ft(x)f_{t}(x) of particles (cells) with position x2x\in{\mathbb{R}}^{2} at time t0t\geq 0 and writes, in the so-called parabolic-elliptic case,

(7) tft(x)+θdivx((Kft)(x)ft(x))=12Δxft(x),whereK(x)=x|x|2.\partial_{t}f_{t}(x)+\theta{\rm div}_{x}((K\star f_{t})(x)f_{t}(x))=\frac{1}{2}\Delta_{x}f_{t}(x),\quad\hbox{where}\quad K(x)=-\frac{x}{|x|^{2}}.

Informally, this solution should be the mean-field limit of the particle system (1) as NN\to\infty.

We refer to the recent review paper on (7) by Arumugam-Tyagi [1]. The best existence of a global solution to (7), including all the subcritical parameters θ(0,2)\theta\in(0,2), is due to Blanchet-Dolbeault-Perthame [2]. The blow-up of solutions to (7), in the supercritical case θ>2\theta>2, has been studied e.g. by Fatkullin [7] and Velasquez [24, 25]. More close to our study, Suzuki [23] has shown, still in the supercritical case, the appearance of a Dirac mass with a precise (critical) weight, at explosion. This is the equivalent, in the limit NN\to\infty, to the fact that limtζXt\lim_{t\to\zeta-}X_{t} exists and corresponds to a KK-collision, for some K[[1,N]]K\subset[\![1,N]\!] with precise cardinal k0k_{0}. Let us finally mention Dolbeault-Schmeiser [6], who propose a post-explosion model in the supercritical case. Concerning particle systems associated with (7), let us mention Stevens [22], who studies a physically more complete particle system with two types of particles, for cells and chemo-attractant particles, with a regularized attraction kernel. Haškovec and Schmeiser [12, 13] study a particle system closer to (1), but with, again, a regularized attraction kernel. Cattiaux-Pédèches [4], as well as [8], study the system (1) without regularization in the subcritical case: existence of a global solution to (1) has been shown in [8] when θ(0,2(N2)/(N1))\theta\in(0,2(N-2)/(N-1)), and uniqueness of this solution has been established in [4]. Also, the theory of Dirichlet spaces has been used in [4] to build a solution to (1). Finally, the limit as NN\to\infty to a solution of (7) is proved in [8] in the very subcritical case where θ(0,1/2)\theta\in(0,1/2), up to extraction of a subsequence. This last result has been improved by Bresch-Jabin-Wang [3], who remove the necessity of extracting a subsequence and consider the (still very subcritical) case where θ(0,1)\theta\in(0,1). Olivera-Richard-Tomasevic [18] have recently established the NN\to\infty convergence of a smoothed version of (1), for all the subcritical cases θ(0,2)\theta\in(0,2). Informally, in view of the mean distance between particles, the regularization used in [18] is not far from being physically reasonable. There is also a related paper of Jabir-Talay-Tomasevic [14] about a one-dimensional but more complicated parabolic-parabolic model.

Let us finally mention the seminal paper of Osada [19], see also [9] for a more recent study, which concerns the vortex model: this is very close to (1), but the attraction x/|x|2-x/|x|^{2} is replaced by a rotating interaction x/|x|2x^{\perp}/|x|^{2}, so that particles never encounter.

1.10. Originality and difficulties

To our knowledge, this is the first study of the supercritical Keller-Segel particle system near explosion. We hope that this model, which makes compete diffusion and Coulomb interactions, is very natural from a physical point of view, beyond the Keller-Segel community. The phenomenon we discovered seems surprising and original, in particular because of the gap between binary and k2k_{2}-ary collisions. We are not aware of other works, possibly dealing with other models, showing such a behavior.

In Section 3, we give the main arguments of the proofs, with quite a high level of precision, but ignoring the technical issues. While it is rather clear, intuitively, that the process explodes in finite time when θ2\theta\geq 2 and that no KK-collisions may occur for |K|[[3,k21]]|K|\in[\![3,k_{2}-1]\!], the continuity at explosion is delicate, and some rather deep arguments are required to show that that each k2k_{2}-ary collision is preceded by many binary collisions, that each k1k_{1}-ary collision is preceded by many k2k_{2}-ary collisions, that explosion is preceded by many k1k_{1}-ary collisions, and that explosion is due to the emergence of a cluster with precise size k0k_{0} (which more or less says that a possible (k0+1)(k_{0}+1)-ary collision would necessarily be preceded by a k0k_{0}-collision).

Actually, the rigorous proofs are made technically much more involved than those presented in Section 3, because we have to use the theory of Dirichlet spaces. Due to the singularity of the interactions and to the occurrence of many collisions near explosion, we can unfortunately not, as already mentioned, deal at the rigorous level directly with the S.D.E. (1). We thus have to use some suitable heavy versions of some usual tools such as Itô’s formula, Girsanov’s theorem, time-change, etc.

1.11. Plan of the paper

In Section 2, we introduce some notation of constant use. In Section 3, we explain the main ideas of the proofs, with quite a high level of precision, but without speaking of the heavy technical issues related to the use of the theory of Dirichlet spaces. Section 4 is devoted to the existence of a first version of the Keller-Segel process, namely without the property that xXXt1{\mathbb{P}}^{X}_{x}\circ X_{t}^{-1} has a density, and we introduce a spherical Keller-Segel process. In Section 5, we show that the Keller-Segel process enjoys a crucial and noticeable decomposition in terms of a 22-dimensional Brownian motion, a squared Bessel process and a spherical process. Section 6 consists in building some smooth approximations of some indicator functions that behave well under the action of the generator X{\mathcal{L}}^{X}. In Section 7, we make use of the Girsanov theorem to prove that when two sets of particles of a KSKS-process are not too close from each other, they behave as two independent smaller KSKS-processes. In Section 8, we study explosion and continuity (in the usual sense) at the explosion time. Section 9 is devoted to establish some parts of Theorem 1.5 for some particular ranges of values of NN and θ\theta. Using the results of Section 7, we reduce the general study to the special cases of Section 9 and we prove, in Section 10, that the conclusions of Theorem 1.5 hold true quasi-everywhere. Finally, in Section 11, we remove the restriction quasi-everywhere and conclude the proofs of Propositions 1.2 and 1.3 and of Theorem 1.5. Appendix A contains a few elementary computations: proof of Lemma 1.1, proof that μ\mu is Radon on Ek0E_{k_{0}}, and study of a similar measure on a sphere. We end the paper with Appendix B, that summarizes all the notions and results about Dirichlet spaces and Hunt processes we shall use.

2. Notation

We introduce the spaces

H={x(2)N:S[[1,N]](x)=0},S={x(2)N:i=1Nxi2=1}and𝕊=HS.\displaystyle H=\Big{\{}x\in({\mathbb{R}}^{2})^{N}:S_{[\![1,N]\!]}(x)=0\Big{\}},\quad S=\Big{\{}x\in({\mathbb{R}}^{2})^{N}:\sum_{i=1}^{N}\|x^{i}\|^{2}=1\Big{\}}\quad\hbox{and}\quad{\mathbb{S}}=H\cap S.

For u𝕊u\in{\mathbb{S}}, we have S[[1,N]](u)=0S_{[\![1,N]\!]}(u)=0 and R[[1,N]](u)=1R_{[\![1,N]\!]}(u)=1. We consider the (unnormalized) Lebesgue measure σ\sigma on 𝕊{\mathbb{S}}, as well as, recall (4),

(1) β(du)=𝐦(u)σ(du).\beta({\rm d}u)={\mathbf{m}}(u)\sigma({\rm d}u).

We define γ:2(2)N\gamma:{\mathbb{R}}^{2}\to({\mathbb{R}}^{2})^{N} by γ(z)=(z,,z)\gamma(z)=(z,\dots,z) and Ψ:2×+×𝕊EN(2)N\Psi:{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}^{*}\times{\mathbb{S}}\to E_{N}\subset({\mathbb{R}}^{2})^{N} by

(2) Ψ(z,r,u)=γ(z)+ru,i.e.(Ψ(z,r,u))i=zruifori[[1,N]].\Psi(z,r,u)=\gamma(z)+\sqrt{r}\;u,\qquad\hbox{i.e.}\qquad(\Psi(z,r,u))^{i}=z-\sqrt{r}u^{i}\quad\hbox{for}\quad i\in[\![1,N]\!].

We have S[[1,N]](Ψ(z,r,u))=zS_{[\![1,N]\!]}(\Psi(z,r,u))=z and R[[1,N]](Ψ(z,r,u))=rR_{[\![1,N]\!]}(\Psi(z,r,u))=r. The orthogonal projection πH:(2)NH\pi_{H}:({\mathbb{R}}^{2})^{N}\to H is given by

πH(x)=xγ(S[[1,N]](x)),i.e.(πH(x))i=xiS[[1,N]](x)fori[[1,N]]\pi_{H}(x)=x-\gamma(S_{[\![1,N]\!]}(x)),\qquad\hbox{i.e.}\qquad(\pi_{H}(x))^{i}=x^{i}-S_{[\![1,N]\!]}(x)\quad\hbox{for}\quad i\in[\![1,N]\!]

and we introduce Φ𝕊:EN𝕊\Phi_{\mathbb{S}}:E_{N}\to{\mathbb{S}} defined by

(3) Φ𝕊(x)=πHxπHx,i.e.(Φ𝕊(x))i=xiS[[1,N]](x)R[[1,N]](x)fori[[1,N]].\displaystyle\Phi_{\mathbb{S}}(x)=\frac{\pi_{H}x}{||\pi_{H}x||},\qquad\hbox{i.e.}\qquad(\Phi_{\mathbb{S}}(x))^{i}=\frac{x^{i}-S_{[\![1,N]\!]}(x)}{\sqrt{R_{[\![1,N]\!]}(x)}}\quad\hbox{for}\quad i\in[\![1,N]\!].

For x(2)N{0}x\in({\mathbb{R}}^{2})^{N}\setminus\{0\}, the projections πx:(2)Nx\pi_{x^{\perp}}:({\mathbb{R}}^{2})^{N}\to x^{\perp} and πx:(2)N\pi_{x}:({\mathbb{R}}^{2})^{N}\to span(x)(x) are given by

πx(y)=yxyx2xandπx(y)=xyx2x,\pi_{x^{\perp}}(y)=y-\frac{x\cdot y}{||x||^{2}}x\qquad\hbox{and}\qquad\pi_{x}(y)=\frac{x\cdot y}{||x||^{2}}x,

where xy=i=1Nxiyix\cdot y=\sum_{i=1}^{N}x^{i}\cdot y^{i}.

We denote by b:E2(2)Nb:E_{2}\to({\mathbb{R}}^{2})^{N} the drift coefficient of (1): for x=(x1,,xN)E2x=(x^{1},\dots,x^{N})\in E_{2},

(4) b(x)=𝐦(x)2𝐦(x)=log𝐦(x)2(2)N,i.e.bi(x)=θNjixixjxixj22b(x)=\frac{\nabla{\mathbf{m}}(x)}{2{\mathbf{m}}(x)}\color[rgb]{0,0,0}=\frac{\nabla\log{\mathbf{m}}(x)}{2}\color[rgb]{0,0,0}\in({\mathbb{R}}^{2})^{N},\qquad\hbox{i.e.}\qquad b^{i}(x)=-\frac{\theta}{N}\sum_{j\neq i}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}}\in{\mathbb{R}}^{2}

for i[[1,N]]i\in[\![1,N]\!]. Indeed, since log𝐦(x)=θ2N1ijNlogxixj2\log{\mathbf{m}}(x)=-\frac{\theta}{2N}\sum_{1\leq i\neq j\leq N}\log||x^{i}-x^{j}||^{2}, we e.g. have

x1log𝐦(x)2=θ4Nx1[i=2Nlogxix12+j=2Nlogx1xj2]=θ2Nx1j=2Nlogx1xj2,\frac{\nabla_{x^{1}}\log{\mathbf{m}}(x)}{2}=-\frac{\theta}{4N}\nabla_{x^{1}}\Big{[}\sum_{i=2}^{N}\log||x^{i}-x^{1}||^{2}+\sum_{j=2}^{N}\log||x^{1}-x^{j}||^{2}\Big{]}=-\frac{\theta}{2N}\nabla_{x^{1}}\sum_{j=2}^{N}\log||x^{1}-x^{j}||^{2},

whence

x1log𝐦(x)2=θNj=2Nx1xjx1xj2.\frac{\nabla_{x^{1}}\log{\mathbf{m}}(x)}{2}=-\frac{\theta}{N}\sum_{j=2}^{N}\frac{x^{1}-x^{j}}{\|x^{1}-x^{j}\|^{2}}.

Finally, we introduce the natural operators defined for φC1(𝕊)\varphi\in C^{1}({\mathbb{S}}) and u𝕊u\in{\mathbb{S}} by

(5) 𝕊φ(u)=[φΦ𝕊](u)(2)NandΔ𝕊φ(u)=Δ[φΦ𝕊](u),\nabla_{\mathbb{S}}\varphi(u)=\nabla[\varphi\circ\Phi_{\mathbb{S}}](u)\in({\mathbb{R}}^{2})^{N}\quad\hbox{and}\quad\Delta_{\mathbb{S}}\varphi(u)=\Delta[\varphi\circ\Phi_{\mathbb{S}}](u)\in{\mathbb{R}},

where \nabla and Δ\Delta stand for the usual gradient and Laplacian in (2)N({\mathbb{R}}^{2})^{N}. Since 𝕊EN(2)N{\mathbb{S}}\subset E_{N}\subset({\mathbb{R}}^{2})^{N}, with ENE_{N} open, and since Φ𝕊\Phi_{\mathbb{S}} is smooth on ENE_{N}, we can indeed define [φΦ𝕊](u)\nabla[\varphi\circ\Phi_{\mathbb{S}}](u) and Δ[φΦ𝕊](u)\Delta[\varphi\circ\Phi_{\mathbb{S}}](u) for all u𝕊u\in{\mathbb{S}}. Similarly, for φC1(𝕊,(2)N)\varphi\in C^{1}({\mathbb{S}},({\mathbb{R}}^{2})^{N}) and u𝕊u\in{\mathbb{S}}, we set

(6) div𝕊φ(u)=div[φΦ𝕊](u).{\rm div}_{\mathbb{S}}\varphi(u)={\rm div}[\varphi\circ\Phi_{\mathbb{S}}](u)\in{\mathbb{R}}.

To conclude this subsection, we note that for all φC((2)N)\varphi\in C^{\infty}(({\mathbb{R}}^{2})^{N}), for all u𝕊u\in{\mathbb{S}},

(7) 𝕊(φ|𝕊)(u)=πH(πu(φ(u))).\displaystyle\nabla_{\mathbb{S}}(\varphi|_{\mathbb{S}})(u)=\pi_{H}(\pi_{u^{\perp}}(\nabla\varphi(u))).

Indeed, it suffices to observe that setting G(x)=x/xG(x)=x/||x|| for all x(2)N{0}x\in({\mathbb{R}}^{2})^{N}\setminus\{0\}, we have Φ𝕊=GπH\Phi_{\mathbb{S}}=G\circ\pi_{H}, dxG=πx/x{\rm d}_{x}G=\pi_{x^{\perp}}/||x|| and dxπH=πH{\rm d}_{x}\pi_{H}=\pi_{H} and that for u𝕊u\in{\mathbb{S}}, we have πH(u)=u\pi_{H}(u)=u and πH(u)=1||\pi_{H}(u)||=1.

3. Main ideas of the proofs

Here we explain the main ideas of the proofs of Proposition 1.3 and Theorem 1.5. The arguments below are completely informal. In particular, we do as if our KS(θ,N)KS(\theta,N)-process (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta)} was a true solution to (1) until explosion and we apply Itô’s formula without care. We always assume at least that N2N\geq 2, θ>0\theta>0 and N>θN>\theta, which implies that k0=2N/θ3k_{0}=\lceil 2N/\theta\rceil\geq 3.

3.1. Existence

The existence of the KS(θ,N)KS(\theta,N)-process (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta)}, with values in Ek0E_{k_{0}}, is an easy application of Fukushima-Oshima-Takeda [11, Theorem 7.2.1]. The only difficulty is to show that the invariant measure μ\mu is a Radon on Ek0E_{k_{0}}, see Proposition A.1. The process may explode, i.e. get out of any compact subset of Ek0E_{k_{0}} in finite time. Observe that a typical compact subset of Ek0E_{k_{0}} is of the form, for ε>0\varepsilon>0,

𝒦ε={x(2)N:x1/ε and for all K[[1,N]] such that |K|=k0,RK(x)ε}.{\mathcal{K}}_{\varepsilon}=\{x\in({\mathbb{R}}^{2})^{N}:||x||\leq 1/\varepsilon\mbox{ and for all }K\subset[\![1,N]\!]\mbox{ such that }|K|=k_{0},\;R_{K}(x)\geq\varepsilon\}.

3.2. Center of mass and dispersion process

One can verify, using Itô’s formula, that the center of mass S[[1,N]](X)S_{[\![1,N]\!]}(X) is a 22-dimensional Brownian motion with diffusion constant N1/2N^{-1/2}, that the dispersion process R[[1,N]](X)R_{[\![1,N]\!]}(X) is a squared Bessel process with dimension dθ,N(N)d_{\theta,N}(N), recall (2), and that these two processes are independent.

Consequently, if ζ<\zeta<\infty, the limits limtζS[[1,N]](Xt)\lim_{t\to\zeta-}S_{[\![1,N]\!]}(X_{t}) and limtζR[[1,N]](Xt)\lim_{t\to\zeta-}R_{[\![1,N]\!]}(X_{t}) a.s. exist, and this implies that lim suptζXt<\limsup_{t\to\zeta-}||X_{t}||<\infty: the process cannot explode to infinity, it can only explode because it tends to the boundary of Ek0E_{k_{0}}. If moreover k0>Nk_{0}>N (i.e. if θ<2\theta<2), this is sufficient to show that ζ=\zeta=\infty, since then Ek0=(2)NE_{k_{0}}=({\mathbb{R}}^{2})^{N}.

3.3. Behavior of distant subsets of particles

Consider a partition K1,,KpK_{1},\dots,K_{p} of [[1,N]][\![1,N]\!]. If we neglect interactions between particles of which the indexes are not in the same subset, we have, for each [[1,p]]\ell\in[\![1,p]\!], setting θ~=θ|K|/N{\tilde{\theta}}_{\ell}=\theta|K_{\ell}|/N,

dXti=dBtiθ~|K|jK{i}XtiXtjXtiXtj2dt,iK{\rm d}X^{i}_{t}={\rm d}B^{i}_{t}-\frac{{\tilde{\theta}}_{\ell}}{|K_{\ell}|}\sum_{j\in K_{\ell}\setminus\{i\}}\frac{X^{i}_{t}-X^{j}_{t}}{\|X^{i}_{t}-X^{j}_{t}\|^{2}}{\rm d}t,\qquad i\in K_{\ell}

and we recognize a KS(θ~,|K|)KS({\tilde{\theta}}_{\ell},|K_{\ell}|)-process.

During time intervals where particles indexed in different subsets are far enough from each other, we can indeed bound the interaction between those particles, so that the Girsanov theorem tells us that (Xti)iK1,,(Xti)iKp(X^{i}_{t})_{i\in K_{1}},\dots,(X^{i}_{t})_{i\in K_{p}} behave similarly, in the sense of trajectories, as independent KS(θ~1,|K1|)KS({\tilde{\theta}}_{1},|K_{1}|), …, KS(θ~p,|Kp|)KS({\tilde{\theta}}_{p},|K_{p}|)-processes.

3.4. Brownian and Bessel behaviors of isolated subsets of particles

Consider K[[1,N]]K\subset[\![1,N]\!]. As seen just above, during time intervals where the particles indexed in KK are far from all the other ones, the system (Xti)iK(X^{i}_{t})_{i\in K} behaves, in the sense of trajectories, like a KS(θ|K|/N,|K|)KS(\theta|K|/N,|K|)-process. Hence, as seen in Subsection 3.2, SK(Xt)S_{K}(X_{t}) behaves like a 22-dimensional Brownian motion with diffusion constant |K|1/2|K|^{-1/2} and RK(Xt)R_{K}(X_{t}) behaves like a squared Bessel process of dimension dθ|K|/N,|K|(|K|)d_{\theta|K|/N,|K|}(|K|), which equals dθ,N(|K|)d_{\theta,N}(|K|), recall (2).

3.5. Continuity at explosion

Here we assume that N>θ2N>\theta\geq 2, so that k0[[2,N]]k_{0}\in[\![2,N]\!] and we explain why a.s., ζ<\zeta<\infty and Xζ=limtζXtX_{\zeta-}=\lim_{t\to\zeta-}X_{t} exists, in the usual sense of (2)N({\mathbb{R}}^{2})^{N}.

(a) We first show that ζ<\zeta<\infty a.s. On the event where ζ=\zeta=\infty, the squared Bessel process R[[1,N]](X)R_{[\![1,N]\!]}(X) is defined for all times. Recall that dθ,N(N)0d_{\theta,N}(N)\leq 0 (because θ2\theta\geq 2) and that a squared Bessel process with negative dimension can be defined on the whole time half-line and a.s. becomes negative in finite time. Since R[[1,N]](X)0R_{[\![1,N]\!]}(X)\geq 0 by definition, this contradicts the fact that ζ=\zeta=\infty. Similarly, one can show that a KS(θ,N)KS(\theta,N)-process has no chance to be defined after the first hitting time τK\tau_{K} of 0 by RK(Xt)R_{K}(X_{t}), where |K|=k0|K|=k_{0}: this makes the choice of the space Ek0E_{k_{0}} very natural. Indeed, assume that XX is defined during [0,ζ)[0,\zeta^{\prime}) with ζ>τK\zeta^{\prime}>\tau_{K}. Consider the maximal subset LL of [[1,N]][\![1,N]\!] containing KK and such that RL(XτK)=0R_{L}(X_{\tau_{K}})=0. Then there is ε>0\varepsilon>0 such that during [τK,τK+ε][0,ζ)[\tau_{K},\tau_{K}+\varepsilon]\subset[0,\zeta^{\prime}), the particles labeled in LL are far from the ones labeled outside LL. By Subsection 3.4, (RL(XτK+t))t[0,ε](R_{L}(X_{\tau_{K}+t}))_{t\in[0,\varepsilon]} behaves like a squared Bessel process with dimension dθ,N(|L|)d_{\theta,N}(|L|) issued from 0. But such a process is instantaneously negative, because dθ,N(|L|)0d_{\theta,N}(|L|)\leq 0 (since |L|k0|L|\geq k_{0}). Since RL(X)0R_{L}(X)\geq 0, this contradicts the fact that τK[0,ζ)\tau_{K}\in[0,\zeta^{\prime}).

(b) We next show by reverse induction that a.s. for all K[[1,N]]K\subset[\![1,N]\!] with |K|2|K|\geq 2, we have

(1) eitherlimtζRK(Xt)=0orlim inftζRK(Xt)>0.\mbox{either}\quad\lim_{t\to\zeta-}R_{K}(X_{t})=0\quad\mbox{or}\quad\liminf_{t\to\zeta-}R_{K}(X_{t})>0.

If K=[[1,N]]K=[\![1,N]\!], limtζRK(Xt)\lim_{t\to\zeta-}R_{K}(X_{t}) exists by continuity of the (true) squared Bessel process RK(Xt)R_{K}(X_{t}) and this implies the result. We now fix n[[3,N]]n\in[\![3,N]\!] and assume that (1) holds true for all KK such that |K|n|K|\geq n. We consider K[[1,N]]K\subset[\![1,N]\!] with |K|=n1|K|=n-1: by induction assumption, either there is iKi\notin K such that limtζRK{i}(Xt)=0\lim_{t\to\zeta-}R_{K\cup\{i\}}(X_{t})=0 and then limtζRK(Xt)=0\lim_{t\to\zeta-}R_{K}(X_{t})=0, or for all i[[1,N]]Ki\in[\![1,N]\!]\setminus K, lim inftζRK{i}(Xt)>0\liminf_{t\to\zeta-}R_{K\cup\{i\}}(X_{t})>0. In this last case, and when lim suptζRK(Xt)>0\limsup_{t\to\zeta-}R_{K}(X_{t})>0 and lim inftζRK(Xt)=0\liminf_{t\to\zeta-}R_{K}(X_{t})=0 (which is the negation of (1)), there are α>0\alpha>0 and ε>0\varepsilon>0 such that (i) RK(Xt)R_{K}(X_{t}) upcrosses [ε/2,ε][\varepsilon/2,\varepsilon] infinitely often during [ζα,ζ)[\zeta-\alpha,\zeta) and (ii) for all t[ζα,ζ)t\in[\zeta-\alpha,\zeta) such that RK(Xt)<εR_{K}(X_{t})<\varepsilon, the particles indexed in KK are far from all the other ones (because then RK(Xt)R_{K}(X_{t}) is small and RK{i}(Xt)R_{K\cup\{i\}}(X_{t}) is large for all iKi\notin K), so that RK(Xt)R_{K}(X_{t}) behaves like a squared Bessel process with dimension dθ,N(|K|)d_{\theta,N}(|K|), see Subsection 3.4. Points (i) and (ii) are in contradiction, since a squared Bessel process is continuous and thus cannot upcross [ε/2,ε][\varepsilon/2,\varepsilon] infinitely often during a finite time interval.

(c) We now show that limtζXt\lim_{t\to\zeta-}X_{t} exists. Using (b) and the (random) equivalence relation on [[1,N]][\![1,N]\!] defined by iji\sim j if and only if limtζR{i,j}(Xt)=0\lim_{t\to\zeta-}R_{\{i,j\}}(X_{t})=0, one can build a (random) partition 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} of [[1,N]][\![1,N]\!] such that for all p[[1,]]p\in[\![1,\ell]\!], limtζRKp(Xt)=0\lim_{t\to\zeta-}R_{K_{p}}(X_{t})=0 and lim inftζminiKpRKp{i}(Xt)>0\liminf_{t\to\zeta-}\min_{i\notin K_{p}}R_{K_{p}\cup\{i\}}(X_{t})>0. Hence, there is α[0,ζ)\alpha\in[0,\zeta) such that for all pqp\neq q, the particles labeled in KpK_{p} are far from the ones labeled in KqK_{q} during [α,ζ)[\alpha,\zeta). As seen in Subsection 3.4, we conclude that for all p[[1,]]p\in[\![1,\ell]\!], SKp(Xt)S_{K_{p}}(X_{t}) behaves like a Brownian motion during [α,ζ)[\alpha,\zeta), and thus Mp=limtζSKp(Xt)M_{p}=\lim_{t\to\zeta-}S_{K_{p}}(X_{t}) exists. Since moreover limtζRKp(Xt)=0\lim_{t\to\zeta-}R_{K_{p}}(X_{t})=0, we deduce that for all iKpi\in K_{p}, limtζXti=Mp\lim_{t\to\zeta-}X^{i}_{t}=M_{p}. As a conclusion limtζXti\lim_{t\to\zeta-}X^{i}_{t} exists for all i[[1,N]]i\in[\![1,N]\!].

3.6. A spherical process

We recall that 𝕊{\mathbb{S}}, πH\pi_{H}, πu\pi_{u^{\perp}} and bb were introduced in Section 2 and introduce the possibly exploding (with life-time ξ\xi) process (Ut)t[0,ξ)(U_{t})_{t\in[0,\xi)} with values in 𝕊Ek0{\mathbb{S}}\cap E_{k_{0}}, informally solving (we will also use here the theory of Dirichlet spaces), for some given U0𝕊Ek0U_{0}\in{\mathbb{S}}\cap E_{k_{0}} and some (2)N({\mathbb{R}}^{2})^{N}-valued Brownian motion (Bt)t0(B_{t})_{t\geq 0},

Ut=U0+0tπUsπHdBs+0tπUsπHb(Us)ds2N320tUsds.U_{t}=U_{0}+\int_{0}^{t}\pi_{U_{s}^{\perp}}\pi_{H}{\rm d}B_{s}+\int_{0}^{t}\pi_{U_{s}^{\perp}}\pi_{H}b(U_{s}){\rm d}s-\frac{2N-3}{2}\int_{0}^{t}U_{s}{\rm d}s.

We call such a process a SKS(θ,N)SKS(\theta,N)-process.

One can check that this process is β\beta-symmetric, where β\beta is defined in (1), and that β\beta is Radon on 𝕊Ek0{\mathbb{S}}\cap E_{k_{0}}, see Proposition A.3. And we will see that if k0Nk_{0}\geq N, then β(𝕊)<\beta({\mathbb{S}})<\infty, so that the process (Ut)t0(U_{t})_{t\geq 0} is non-exploding and positive recurrent.

3.7. Decomposition of the process

We assume that N2N\geq 2 and θ>0\theta>0 are such dθ,N(N)<2d_{\theta,N}(N)<2 and, as usual, N>θN>\theta. We consider a 22-dimensional Brownian (Mt)t0(M_{t})_{t\geq 0} with diffusion constant N1/2N^{-1/2}, a squared Bessel process (Dt)t[0,τD)(D_{t})_{t\in[0,\tau_{D})} with dimension dθ,N(N)d_{\theta,N}(N) killed when it hits 0, with life-time τD\tau_{D}, and a SKS(θ,N)SKS(\theta,N)-process (Ut)t[0,ξ)(U_{t})_{t\in[0,\xi)}, these three processes being independent. We introduce the time-change

At=0tdsDs,t[0,τD).A_{t}=\int_{0}^{t}\frac{{\rm d}s}{D_{s}},\quad t\in[0,\tau_{D}).

Since τD<\tau_{D}<\infty (because dθ,N(N)<2d_{\theta,N}(N)<2), since DτD=0D_{\tau_{D}}=0 and since, roughly, the paths of (Dt)t[0,τD)(\sqrt{D_{t}})_{t\in[0,\tau_{D})} are 1/21/2-Hölder continuous, it holds that AτD=A_{\tau_{D}}=\infty a.s. We introduce the inverse function ρ:[0,)[0,τD)\rho:[0,\infty)\to[0,\tau_{D}) of A:[0,τD)[0,)A:[0,\tau_{D})\to[0,\infty).

We also set ζ=ρξ\zeta^{\prime}=\rho_{\xi} and observe that ζτD\zeta^{\prime}\leq\tau_{D}, since ρ\rho is [0,τD)[0,\tau_{D})-valued, and that ζ<τD\zeta^{\prime}<\tau_{D} if and only if ξ<\xi<\infty. A fastidious but straightforward computation shows that, recalling (2),

Xt=Ψ(Mt,Dt,UAt),i.e.Xti=Mt+DtUAti,i[[1,N]],X_{t}=\Psi(M_{t},D_{t},U_{A_{t}}),\qquad\hbox{i.e.}\quad X_{t}^{i}=M_{t}+\sqrt{D_{t}}U_{A_{t}}^{i},\qquad i\in[\![1,N]\!],

which is well-defined during [0,ζ)[0,\zeta^{\prime}), solves (1).

This decomposition of the KS(θ,N)KS(\theta,N)-process, which is noticeable in that UU satisfies an autonomous S.D.E. and thus is Markov, is at the basis of our analysis.

In other words, (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta^{\prime})} is the restriction to the time interval [0,ζ)[0,\zeta^{\prime}) of a KS(θ,N)KS(\theta,N)-process (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta)}. Moreover, we have ζ=ζτD\zeta^{\prime}=\zeta\land\tau_{D}: if ξ\xi is finite, then UU gets out of 𝕊Ek0{\mathbb{S}}\cap E_{k_{0}} at time ξ\xi, so that XX gets out of Ek0E_{k_{0}} at time ζ=ρξ<τD\zeta^{\prime}=\rho_{\xi}<\tau_{D}, whence ζ=ζ=ζτD\zeta=\zeta^{\prime}=\zeta\land\tau_{D}; if next ξ=\xi=\infty, then ζ=τD\zeta^{\prime}=\tau_{D} and UU remains in Ek0E_{k_{0}} for all times, so that XX remains in Ek0E_{k_{0}} during [0,τD)[0,\tau_{D}), whence ζτD\zeta\geq\tau_{D}.

We have S[[1,N]](Xt)=MtS_{[\![1,N]\!]}(X_{t})=M_{t} and R[[1,N]](Xt)=DtR_{[\![1,N]\!]}(X_{t})=D_{t} for all t[0,ζτD)t\in[0,\zeta\land\tau_{D}), because UU is 𝕊{\mathbb{S}}-valued. By definition of 𝕊{\mathbb{S}}, the process UU cannot have any [[1,N]][\![1,N]\!]-collision. But for any K[[1,N]]K\subset[\![1,N]\!] with cardinal at most N1N-1,

(2) U has a K-collision at t[0,ξ) if and only if X has a K-collision at ρt[0,ζτD).\displaystyle\hbox{$U$ has a $K$-collision at $t\in[0,\xi)$ if and only if $X$ has a $K$-collision at $\rho_{t}\in[0,\zeta\land\tau_{D})$}.

Moreover, as seen a few lines above, ξ<\xi<\infty is equivalent to ζ<τD\zeta<\tau_{D}. In other words, since R[[1,N]](Xt)=DtR_{[\![1,N]\!]}(X_{t})=D_{t} for all t[0,ζτD)t\in[0,\zeta\land\tau_{D}) and since τD=inf{t>0:Dt=0}\tau_{D}=\inf\{t>0:D_{t}=0\}, we have

(3) ξ<if and only ifinft[0,ζ)R[[1,N]](Xt)>0.\displaystyle\xi<\infty\qquad\hbox{if and only if}\qquad\inf_{t\in[0,\zeta)}R_{[\![1,N]\!]}(X_{t})>0.

3.8. Some special cases

Using the Girsanov theorem, see Subsection 3.4, we will manage to reduce a large part of the study to the special cases that we examine in the present subsection. Here we explain the following facts, for N2N\geq 2 and θ>0\theta>0 with N>θN>\theta: (a) if dθ,N(N1)(0,2)d_{\theta,N}(N-1)\in(0,2), then a.s., τD=inf{t>0:R[[1,N]](Xt)=0}ζ\tau_{D}=\inf\{t>0:R_{[\![1,N]\!]}(X_{t})=0\}\leq\zeta and for all r[0,τD)r\in[0,\tau_{D}), all K[[1,N]]K\subset[\![1,N]\!] with |K|=N1|K|=N-1, (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta)} has infinitely many KK-collisions during [r,τD)[r,\tau_{D}); (b) if dθ,N(N1)0d_{\theta,N}(N-1)\leq 0 (whence k0N1k_{0}\leq N-1), then a.s., inft[0,ζ)R[[1,N]](Xt)>0\inf_{t\in[0,\zeta)}R_{[\![1,N]\!]}(X_{t})>0.

We keep the same notation as in the previous subsection.

(i) We first verify that in (a), τDζ\tau_{D}\leq\zeta. Since dθ,N(N1)(0,2)d_{\theta,N}(N-1)\in(0,2), it holds that k0Nk_{0}\geq N. If first k0>Nk_{0}>N, then ζ=\zeta=\infty by Subsection 3.2 and we are done. If next k0=Nk_{0}=N, then ζ<\zeta<\infty and XζX_{\zeta-} exists by Subsection 3.5. Moreover XζX_{\zeta-} cannot belong to Ek0=ENE_{k_{0}}=E_{N} by definition of ζ\zeta and thus has its NN particles at the same place, i.e. R[[1,N]](Xζ)=0R_{[\![1,N]\!]}(X_{\zeta-})=0: we have ζ=τD\zeta=\tau_{D}.

(ii) In (b), ζ<\zeta<\infty by Subsection 3.5 because dθ,N(N1)0d_{\theta,N}(N-1)\leq 0 implies that θ2\theta\geq 2.

(iii) We consider, in any case, the spherical process (Ut)t[0,ξ)(U_{t})_{t\in[0,\xi)} and assume that ξ=\xi=\infty. An Itô computation shows that for K[[1,N]]K\subset[\![1,N]\!], for some 11-dimensional Brownian motion (Wt)t0(W_{t})_{t\geq 0},

dRK(Ut)=\displaystyle{\rm d}R_{K}(U_{t})= 2RK(Ut)(1RK(Ut))dWt+dθ,N(|K|)dtdθ,N(N)RK(Ut)dt\displaystyle 2\sqrt{R_{K}(U_{t})(1-R_{K}(U_{t}))}{\rm d}W_{t}+d_{\theta,N}(|K|){\rm d}t-d_{\theta,N}(N)R_{K}(U_{t}){\rm d}t
2θNiK,jKUtiUtjUtiUtj2(UtiSK(Ut))dt.\displaystyle-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{U^{i}_{t}-U^{j}_{t}}{||U^{i}_{t}-U^{j}_{t}||^{2}}\cdot(U^{i}_{t}-S_{K}(U_{t})){\rm d}t.

We fix ε>0\varepsilon>0 to be chosen later. During time intervals where miniK,jKUtiUtjε\min_{i\in K,j\notin K}\|U^{i}_{t}-U^{j}_{t}\|\geq\varepsilon, we thus have, for some constant CεC_{\varepsilon},

(4) dRK(Ut)\displaystyle{\rm d}R_{K}(U_{t})\leq 2RK(Ut)(1RK(Ut))dWt+dθ,N(|K|)dt+CεRK(Ut)dt,\displaystyle 2\sqrt{R_{K}(U_{t})(1-R_{K}(U_{t}))}{\rm d}W_{t}+d_{\theta,N}(|K|){\rm d}t+C_{\varepsilon}\sqrt{R_{K}(U_{t})}{\rm d}t,

where we used the Cauchy-Schwarz inequality and that RK(Ut)R_{K}(U_{t}) is uniformly bounded (because UU is 𝕊{\mathbb{S}}-valued). Hence, still during time intervals where miniK,jKUtiUtjε\min_{i\in K,j\notin K}\|U^{i}_{t}-U^{j}_{t}\|\geq\varepsilon, by comparison, RK(Ut)R_{K}(U_{t}) is smaller than StS_{t}, the solution to

(5) dSt=2St(1St)dWt+dθ,N(|K|)dt+CεStdt.{\rm d}S_{t}=2\sqrt{S_{t}(1-S_{t})}{\rm d}W_{t}+d_{\theta,N}(|K|){\rm d}t+C_{\varepsilon}\sqrt{S_{t}}{\rm d}t.

And a little study involving scale functions/speed measures shows that this process hits zero in finite time if and only if dθ,N(|K|)<2d_{\theta,N}(|K|)<2, exactly as a squared Bessel process with dimension dθ,N(|K|)d_{\theta,N}(|K|).

(iv) We end the proof of (a). In this case, k0Nk_{0}\geq N, so that UU is non-exploding, as seen in Subsection 3.6. Hence ξ=\xi=\infty and we can use (iii). Moreover, UU is recurrent, still by Subsection 3.6. We fix KK with |K|=N1|K|=N-1 and we choose ε>0\varepsilon>0 small enough so that we have

β({u𝕊:miniK,jKuiujε})>0,\beta\Big{(}\Big{\{}u\in{\mathbb{S}}:\min_{i\in K,j\notin K}\|u^{i}-u^{j}\|\geq\varepsilon\Big{\}}\Big{)}>0,

where β\beta is the invariant measure (1) of UU. Hence the process miniK,jKUtiUtj\min_{i\in K,j\notin K}\|U^{i}_{t}-U^{j}_{t}\| visits the zone (ε,)(\varepsilon,\infty) infinitely often and each time, RK(U)R_{K}(U) has a (uniformly) positive probability to hit 0 by (iii) and since dθ,N(|K|)=dθ,N(N1)<2d_{\theta,N}(|K|)=d_{\theta,N}(N-1)<2. Consequently, for any s>0s>0, (Ut)t0(U_{t})_{t\geq 0} has infinitely many KK-collisions during [s,)[s,\infty). Recalling (2) and that ζτD=τD\zeta\land\tau_{D}=\tau_{D} by (i), we conclude that for any r[0,τD)r\in[0,\tau_{D}), (Xt)t[0,ζ)(X_{t})_{t\in[0,\zeta)} has infinitely many KK-collisions during [r,τD)[r,\tau_{D}).

(v) We finally complete the proof of (b). By (3), it is sufficient to show that ξ<\xi<\infty a.s.

Assume that UU is recurrent (and thus non-exploding). Then we take K=[[2,N]]K=[\![2,N]\!] and apply the same reasoning as in (iv): since dθ,N(|K|)0<2d_{\theta,N}(|K|)\leq 0<2, RK(U)R_{K}(U) hits zero in finite time and this makes UU get out of EN1E_{N-1} and thus explode, since UU is (Ek0𝕊)(E_{k_{0}}\cap{\mathbb{S}})-valued and since k0N1k_{0}\leq N-1. We thus have a contradiction.

Hence UU is transient and it eventually gets out of the compact of Ek0𝕊E_{k_{0}}\cap{\mathbb{S}}

𝒦={u𝕊:K[[1,N]] such that |K|=k0, we have RK(u)ε},{\mathcal{K}}=\{u\in{\mathbb{S}}:\forall K\subset[\![1,N]\!]\mbox{ such that }|K|=k_{0},\mbox{ we have }R_{K}(u)\geq\varepsilon\},

for any fixed ε>0\varepsilon>0. Hence on the event where ξ=\xi=\infty, limtmin|K|=k0RK(Ut)=0\lim_{t\to\infty}\min_{|K|=k_{0}}R_{K}(U_{t})=0 a.s. Recalling now that k0N1k_{0}\leq N-1 and that UU is 𝕊{\mathbb{S}}-valued (whence R[[1,N]](Ut)=1R_{[\![1,N]\!]}(U_{t})=1) we can a.s. find KK with |K|[[k0,N1]]|K|\in[\![k_{0},N-1]\!] such that lim inftRK(Ut)=0\liminf_{t\to\infty}R_{K}(U_{t})=0 but lim inftminiKRK{i}(Ut)>0\liminf_{t\to\infty}\min_{i\notin K}R_{K\cup\{i\}}(U_{t})>0. It is then not too hard to find α>0\alpha>0 and ε>0\varepsilon>0 such that each time RK(Ut)<αR_{K}(U_{t})<\alpha (which often happens), all the particles indexed in KK are far from all the other ones with a distance greater than ε>0\varepsilon>0. We conclude from (iii), since dθ,N(|K|)0d_{\theta,N}(|K|)\leq 0 (because |K|k0|K|\geq k_{0}) that each time RK(Ut)<αR_{K}(U_{t})<\alpha, it has a (uniformly) positive probability to hit zero. On the event ξ=\xi=\infty, this will eventually happen, so that the process UU will have a KK-collision and thus will leave Ek0E_{k_{0}} in finite time. Hence UU will explode, so that ξ<\xi<\infty.

3.9. Size of the cluster

We assume that N>3θ6N>3\theta\geq 6. Hence ζ<\zeta<\infty and XζX_{\zeta-} exists, by Subsection 3.5. Moreover, by definition of ζ\zeta, we know that XζEk0X_{\zeta-}\notin E_{k_{0}}. We want now to show that XζEk0+1X_{\zeta-}\in E_{k_{0}+1}, i.e. that the cluster causing explosion is precisely composed of k0k_{0} particles. If k0=Nk_{0}=N, there is nothing to do, since then Ek0+1=(2)NE_{k_{0}+1}=({\mathbb{R}}^{2})^{N}. Now if k0N1k_{0}\leq N-1, we assume by contradiction, that there is K[[1,N]]K\subset[\![1,N]\!] with |K|k0+1|K|\geq k_{0}+1 such that RK(Xζ)=0R_{K}(X_{\zeta-})=0 and miniKRK{i}(Xζ)>0\min_{i\notin K}R_{K\cup\{i\}}(X_{\zeta-})>0. Then there is α>0\alpha>0 such that during [ζα,ζ)[\zeta-\alpha,\zeta), the particles indexed in KK are far from the other ones, so that (Xti)t[0,ζ),iK(X^{i}_{t})_{t\in[0,\zeta),i\in K} behaves like a KS(θ|K|/N,|K|)KS(\theta|K|/N,|K|)-process by Subsection 3.3. Observe now that dθ|K|/N,|K|(|K|1)=dθ,N(|K|1)0d_{\theta|K|/N,|K|}(|K|-1)=d_{\theta,N}(|K|-1)\leq 0 because |K|1k0|K|-1\geq k_{0} and |K|>θ|K|/N|K|>\theta|K|/N because N>θN>\theta. We thus know from the special case (b) of Subsection 3.8 that inft[ζα,ζ)RK(Xt)>0\inf_{t\in[\zeta-\alpha,\zeta)}R_{K}(X_{t})>0, which contradicts the fact that RK(Xζ)=0R_{K}(X_{\zeta-})=0.

3.10. Collisions before explosion

We fix again N>3θ6N>3\theta\geq 6. We recall that k1=k01k_{1}=k_{0}-1 and we show that there are infinitely many k1k_{1}-ary collisions just before explosion. We know from the previous subsection that there exists K0[[1,N]]K_{0}\subset[\![1,N]\!] such that |K0|=k0|K_{0}|=k_{0} and RK0(Xζ)=0R_{K_{0}}(X_{\zeta-})=0 and miniK0RK0{i}(Xζ)>0\min_{i\notin K_{0}}R_{K_{0}\cup\{i\}}(X_{\zeta-})>0. Then there is α>0\alpha>0 such that during [ζα,ζ)[\zeta-\alpha,\zeta), the particles indexed in K0K_{0} are far from the other ones, so that (Xti)iK0(X^{i}_{t})_{i\in K_{0}} behaves like a KS(θk0/N,k0)KS(\theta k_{0}/N,k_{0})-process by Subsection 3.3. Observe now that dθk0/N,k0(k01)=dθ,N(k01)(0,2)d_{\theta k_{0}/N,k_{0}}(k_{0}-1)=d_{\theta,N}(k_{0}-1)\in(0,2) thanks to Lemma 1.1 and that k0>θk0/Nk_{0}>\theta k_{0}/N because N>θN>\theta. We thus know from the special case (a) of Subsection 3.8 that (Xti)iK0(X^{i}_{t})_{i\in K_{0}} has infinitely many (K0{i})(K_{0}\setminus\{i\})-collisions just before ζ\zeta, for all iK0i\in K_{0}.

When k2=k11k_{2}=k_{1}-1, one can show in the very same way that for all KK with |K|=k1|K|=k_{1}, for all iKi\in K, there are infinitely many (K{i})(K\setminus\{i\})-collisions just before each KK-collision. We may also use Subsection 3.8-(a), since dθk1/N,k1(k11)=dθ,N(k2)(0,2)d_{\theta k_{1}/N,k_{1}}(k_{1}-1)=d_{\theta,N}(k_{2})\in(0,2), see Lemma 1.1.

3.11. Absence of other collisions

We want to show that when N>3θ6N>3\theta\geq 6, for K[[1,N]]K\subset[\![1,N]\!] with |K|[[3,k21]]|K|\in[\![3,k_{2}-1]\!], there is no KK-collision during (0,ζ)(0,\zeta). Suppose by contradiction that there is K[[1,N]]K\subset[\![1,N]\!] with |K|[[3,k21]]|K|\in[\![3,k_{2}-1]\!] and t(0,ζ)t\in(0,\zeta) such that RK(Xt)=0R_{K}(X_{t})=0 and for all iKi\notin K, RK{i}(Xt)>0R_{K\cup\{i\}}(X_{t})>0. Then there is α>0\alpha>0 such that during [tα,t][t-\alpha,t], the particles indexed in KK are far from the other ones, so that RK(Xt)R_{K}(X_{t}) behaves like a squared Bessel process with dimension dθ|K|/N,|K|(|K|)d_{\theta|K|/N,|K|}(|K|), see Subsection 3.4. Since dθ|K|/N,|K|(|K|)=dθ,N(|K|)2d_{\theta|K|/N,|K|}(|K|)=d_{\theta,N}(|K|)\geq 2 because |K|[[3,k21]]|K|\in[\![3,k_{2}-1]\!], see Lemma 1.1, such a Bessel process cannot hit zero, whence a contradiction.

3.12. Binary collisions

We still assume that N>3θ6N>3\theta\geq 6, we suppose that there is a KK-collision for some K[[1,N]]K\subset[\![1,N]\!] such that |K|=k2|K|=k_{2} at some time t(0,ζ)t\in(0,\zeta) and we want to show that there are infinitely many binary collisions just before tt. There is α>0\alpha>0 such that the particles indexed in KK are far from all the other ones during [tα,t][t-\alpha,t], so that Subsection 3.3 tells us that (Xti)iK(X^{i}_{t})_{i\in K} behaves like a KS(θk2/N,k2)KS(\theta k_{2}/N,k_{2})-process. We observe that k25k_{2}\geq 5, that dθk2/N,k2(k21)=dθ,N(k21)2d_{\theta k_{2}/N,k_{2}}(k_{2}-1)=d_{\theta,N}(k_{2}-1)\geq 2 and that dθk2/N,k2(k2)=dθ,N(k2)(0,2)d_{\theta k_{2}/N,k_{2}}(k_{2})=d_{\theta,N}(k_{2})\in(0,2) by Lemma 1.1.

We are reduced to show that a KS(θ,N)KS(\theta,N)-process, that we still denote by (Xti)i[[1,N]],t0(X_{t}^{i})_{i\in[\![1,N]\!],t\geq 0}, such that N5N\geq 5, dθ,N(N1)2d_{\theta,N}(N-1)\geq 2 and dθ,N(N)(0,2)d_{\theta,N}(N)\in(0,2), a.s. has infinitely many binary collisions before the first instant τD\tau_{D} of [[1,N]][\![1,N]\!]-collision. Such a process does not explode, because k0>Nk_{0}>N (since dθ,N(N)>0d_{\theta,N}(N)>0), see Subsection 3.2. Hence using (2) (which is licit since dθ,N(N)<2d_{\theta,N}(N)<2), we only have to show that e.g. U1U^{1} collides infinitely often with U2U^{2} during [0,)[0,\infty).

First, one easily gets convinced that the probability that e.g. X1X^{1} collides with X2X^{2} before τD\tau_{D} is positive, because the probability that all the particles are pairwise far from each other, except X1X^{1} and X2X^{2}, during the time interval [0,1][0,1], is positive. On this kind of event, by Subsection 3.4, R{1,2}(Xt)R_{\{1,2\}}(X_{t}) behaves like a squared Bessel process with dimension dθ,N(2)(0,2)d_{\theta,N}(2)\in(0,2) and thus hits zero during [0,1][0,1] (and thus before τD\tau_{D}) with positive probability.

Using again (2), we conclude that the probability that U1U^{1} collides with U2U^{2} in finite time is positive. Since now UU is positive recurrent, recall Subsection 3.6 and that k0>Nk_{0}>N (because dθ,N(N)>0d_{\theta,N}(N)>0), we conclude that U1U^{1} collides infinitely often with U2U^{2} during [0,)[0,\infty) as desired.

3.13. Non-integrability of the drift term

Here we check that when dθ,N(k1)(0,1)d_{\theta,N}(k_{1})\in(0,1), the S.D.E. (1) cannot have a solution in the classical sense, because the drift term is not integrable in time. More precisely, recall that there is some KK-collision at some time τ\tau strictly before explosion, for some K[[1,N]]K\subset[\![1,N]\!] with cardinal k1k_{1}. We now show that a.s., for a>0a>0,

τaτ+ai=1NjiXsiXsjXsiXsj2ds=,\int_{\tau-a}^{\tau+a}\sum_{i=1}^{N}\Big{\|}\sum_{j\neq i}\frac{X^{i}_{s}-X^{j}_{s}}{||X^{i}_{s}-X^{j}_{s}||^{2}}\Big{\|}{\rm d}s=\infty,

which indeed shows the non-integrability of the drift term. Since τ\tau is an instant of KK-collision, there exists a>0a>0 small enough so that during [τa,τ+a][0,ζ)[\tau-a,\tau+a]\subset[0,\zeta), the particles labeled in KK are far from the particles labeled in KcK^{c}. It clearly suffices to show that Z=Z=\infty a.s., where

Z=τaτ+aiKjK,jiXsiXsjXsiXsj2ds.Z=\int_{\tau-a}^{\tau+a}\sum_{i\in K}\Big{\|}\sum_{j\in K,j\neq i}\frac{X^{i}_{s}-X^{j}_{s}}{||X^{i}_{s}-X^{j}_{s}||^{2}}\Big{\|}{\rm d}s.

But

Z=τaτ+af(Vs)RK(Xs)ds,whereVs=(Vsi)iKis defined byVsi=XsiSK(Xs)RK(Xs),Z=\int_{\tau-a}^{\tau+a}\frac{f(V_{s})}{\sqrt{R_{K}(X_{s})}}{\rm d}s,\quad\hbox{where}\quad V_{s}=(V^{i}_{s})_{i\in K}\quad\hbox{is defined by}\quad V^{i}_{s}=\frac{X_{s}^{i}-S_{K}(X_{s})}{\sqrt{R_{K}(X_{s})}},

so that VsV_{s} a.s. belongs to 𝕊K={(vi)iK(2)|K|:iKvi=0,iKvi2=1}{\mathbb{S}}_{K}=\{(v^{i})_{i\in K}\in({\mathbb{R}}^{2})^{|K|}:\sum_{i\in K}v^{i}=0,\;\sum_{i\in K}||v^{i}||^{2}=1\}, and where

f(v)=iKjK,jivivjvivj2f(v)=\sum_{i\in K}\Big{\|}\sum_{j\in K,j\neq i}\frac{v^{i}-v^{j}}{||v^{i}-v^{j}||^{2}}\Big{\|}

for each v𝕊Kv\in{\mathbb{S}}_{K}. Since the invariant measure 𝐦{\mathbf{m}} of XX satisfies 𝐦(E2c)=0{\mathbf{m}}(E_{2}^{c})=0, it a.s. holds true that XsE2X_{s}\in E_{2} for a.e. s[0,ζ)s\in[0,\zeta) (at least for a.e. initial condition), so that a.s., f(Vs)f(V_{s}) is well-defined for a.e. s[0,ζ)s\in[0,\zeta). We now show that ff is bounded from below on 𝕊K{\mathbb{S}}_{K}. We have

f(v)maxiKjK,jivivjvivj21|K|iKjK,jivivjvivj22.f(v)\geq\max_{i\in K}\Big{\|}\sum_{j\in K,j\neq i}\frac{v^{i}-v^{j}}{||v^{i}-v^{j}||^{2}}\Big{\|}\geq\sqrt{\frac{1}{|K|}\sum_{i\in K}\Big{\|}\sum_{j\in K,j\neq i}\frac{v^{i}-v^{j}}{||v^{i}-v^{j}||^{2}}\Big{\|}^{2}}.

Using now the Cauchy-Schwarz inequality and the fact that iKvi2=1\sum_{i\in K}||v^{i}||^{2}=1, we find that

f(v)1|K|iKjK,jivivjvivj2vi=12|K|i,jK,jivivjvivj2(vivj)=|K|(|K|1)2|K|.f(v)\geq\frac{1}{\sqrt{|K|}}\sum_{i\in K}\sum_{j\in K,j\neq i}\frac{v^{i}-v^{j}}{||v^{i}-v^{j}||^{2}}\cdot v^{i}=\frac{1}{2\sqrt{|K|}}\sum_{i,j\in K,j\neq i}\frac{v^{i}-v^{j}}{||v^{i}-v^{j}||^{2}}\cdot(v^{i}-v^{j})=\frac{|K|(|K|-1)}{2\sqrt{|K|}}.

To conclude that Z=Z=\infty a.s., it remains to verify that τaτ+a[RK(Xs)]1/2ds=\int_{\tau-a}^{\tau+a}[R_{K}(X_{s})]^{-1/2}{\rm d}s=\infty a.s. By Subsection 3.4, RK(X)R_{K}(X) behaves like a squared Bessel process with dimension dθ,N(k1)d_{\theta,N}(k_{1}) during [τa,τ+a][\tau-a,\tau+a]. Since dθ,N(k1)(0,1)d_{\theta,N}(k_{1})\in(0,1) and RK(Xτ)=0R_{K}(X_{\tau})=0, we conclude that indeed, τaτ+a[RK(Xs)]1/2ds=\int_{\tau-a}^{\tau+a}[R_{K}(X_{s})]^{-1/2}{\rm d}s=\infty a.s.: this can be shown by comparison with the 11-dimensional Brownian motion.

4. Construction of the Keller-Segel particle system

The aim of this section is to build a first version of the Keller-Segel particle system using the book of Fukushima-Oshima-Takeda [11]. We also build a 𝕊{\mathbb{S}}-valued process for later use.

Proposition 4.1.

We fix N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta, recall that k0=2N/θk_{0}=\lceil 2N/\theta\rceil and that μ\mu and β\beta were defined in (4) and (1). We set 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} and 𝒳=𝒳{}{\mathcal{X}}_{\triangle}={\mathcal{X}}\cup\{\triangle\}, as well as 𝒰=𝕊Ek0{\mathcal{U}}={\mathbb{S}}\cap E_{k_{0}} and 𝒰=𝒰{}{\mathcal{U}}_{\triangle}={\mathcal{U}}\cup\{\triangle\}, where \triangle is a cemetery point. (i) There exists a unique diffusion 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳)\mathbb{X}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}_{x}^{X})_{x\in{\mathcal{X}_{\triangle}}}) with values in 𝒳{\mathcal{X}}_{\triangle}, which is μ\mu-symmetric, with regular Dirichlet space (X,X)({\mathcal{E}}^{X}\!,\!{\mathcal{F}}^{X}) on L2((2)N,μ)L^{2}(({\mathbb{R}}^{2})^{N}\!,\!\mu) with core Cc(𝒳)C^{\infty}_{c}({\mathcal{X}}) defined by

for allφCc(𝒳),X(φ,φ)=12(2)Nφ2dμ.\hbox{for all}\quad\varphi\in C^{\infty}_{c}({\mathcal{X}}),\quad\mathcal{E}^{X}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\|\nabla\varphi\|^{2}{\rm d}\mu.

We call such a process a QKS(θ,N)QKS(\theta,N)-process and denote by ζ=inf{t0:Xt=}\zeta=\inf\{t\geq 0:X_{t}=\triangle\} its life-time.

(ii) There exists a unique diffusion 𝕌=(ΩU,U,(Ut)t0,(uU)u𝒰)\mathbb{U}=(\Omega^{U},{\mathcal{M}}^{U},(U_{t})_{t\geq 0},({\mathbb{P}}^{U}_{u})_{u\in{\mathcal{U}_{\triangle}}}) with values in 𝒰{\mathcal{U}_{\triangle}}, which is β\beta-symmetric, with regular Dirichlet space (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) on L2(𝕊,β){L^{2}({\mathbb{S}},\beta)} with core Cc(𝒰)C^{\infty}_{c}({\mathcal{U}}) defined by

for allφCc(𝒰),U(φ,φ)=12𝕊𝕊φ2dβ.\hbox{for all}\quad\varphi\in C^{\infty}_{c}({\mathcal{U}}),\quad\mathcal{E}^{U}(\varphi,\varphi)=\frac{1}{2}\int_{{\mathbb{S}}}\|\nabla_{\mathbb{S}}\varphi\|^{2}{\rm d}\beta.

We call such a process a QSKS(θ,N)QSKS(\theta,N) -process and denote by ξ=inf{t0:Ut=}\xi=\inf\{t\geq 0:U_{t}=\triangle\} its life-time.

The proof that we can build a KS(θ,N)KS(\theta,N)-process, i.e. a QKS(θ,N)QKS(\theta,N)-process such that xXXt1{\mathbb{P}}_{x}^{X}\circ X_{t}^{-1} has density for all xE2x\in E_{2} and all t>0t>0 will be handled in Section 11.

We refer to Subsection B.1 for some explanations about the notions used in this proposition: link between a diffusion (i.e. a continuous Hunt process), its generator, semi-group and its Dirichlet space, definition of the one-point compactification topology, i.e. the topology endowing 𝒳{\mathcal{X}}_{\triangle} and 𝒰{\mathcal{U}}_{\triangle}, and about the quasi-everywhere notion. The state \triangle is absorbing, i.e. Xt=X_{t}=\triangle for all tζt\geq\zeta and Ut=U_{t}=\triangle for all tξt\geq\xi.

Remark 4.2.

By definition of the one-point compactification topology, for any increasing sequence of compact subsets (𝒦n)n1({\mathcal{K}}_{n})_{n\geq 1} of 𝒳{\mathcal{X}} such that n1𝒦n=𝒳\cup_{n\geq 1}{\mathcal{K}}_{n}={\mathcal{X}}, ζ=limninf{t0:Xt𝒦n}\zeta=\lim_{n\to\infty}\inf\{t\geq 0:X_{t}\notin{\mathcal{K}}_{n}\}. Similarly, for any increasing sequence of compact subsets (n)n1({\mathcal{L}}_{n})_{n\geq 1} of 𝒰{\mathcal{U}} such that n1n=𝒰\cup_{n\geq 1}{\mathcal{L}}_{n}={\mathcal{U}}, ξ=limninf{t0:Utn}\xi=\lim_{n\to\infty}\inf\{t\geq 0:U_{t}\notin{\mathcal{L}}_{n}\}.

The uniqueness stated e.g. in Proposition 4.1-(i) has to be understood in the following sense, see [11, Theorem 4.2.8 p 167]: if we have another diffusion 𝕐=(ΩY,Y,(Yt)t0,(xY)x𝒳){\mathbb{Y}}=(\Omega^{Y},{\mathcal{M}}^{Y},(Y_{t})_{t\geq 0},({\mathbb{P}}_{x}^{Y})_{x\in{\mathcal{X}}}) enjoying the same properties, then quasi-everywhere, the law of (Yt)t0(Y_{t})_{t\geq 0} under xY{\mathbb{P}}_{x}^{Y} equals the law of (Xt)t0(X_{t})_{t\geq 0} under xX{\mathbb{P}}_{x}^{X}. The quasi-everywhere notion depends on the Hunt process under consideration but, as recalled in Subsection B.1, two Hunt processes with the same Dirichlet space share the same quasi-everywhere notion.

Proof of Proposition 4.1.

We start with (i). We consider the bilinear form X{\mathcal{E}}^{X} on Cc(𝒳)C^{\infty}_{c}({\mathcal{X}}) defined by X(φ,φ)=12(2)Nφ2dμ{\mathcal{E}}^{X}(\varphi,\varphi)\!=\!\color[rgb]{0,0,0}\frac{1}{2}\color[rgb]{0,0,0}\int_{({\mathbb{R}}^{2})^{N}}\!||\nabla\varphi||^{2}{\rm d}\mu. It is well-defined, since μ\mu is Radon on 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} by Proposition A.1.

We first show that it is closable, see [11, page 2], i.e. that if (φn)n1Cc(𝒳)(\varphi_{n})_{n\geq 1}\subset C^{\infty}_{c}({\mathcal{X}}) is such that limnφn=0\lim_{n}\varphi_{n}=0 in L2((2)N,μ)L^{2}(({\mathbb{R}}^{2})^{N},\mu) and limn,mX(φnφm,φnφm)=0\lim_{n,m}{\mathcal{E}}^{X}(\varphi_{n}-\varphi_{m},\varphi_{n}-\varphi_{m})=0, then limnX(φn,φn)=0\lim_{n}{\mathcal{E}}^{X}(\varphi_{n},\varphi_{n})=0: since φn\nabla\varphi_{n} is a Cauchy sequence in L2((2)N,μ)L^{2}(({\mathbb{R}}^{2})^{N},\mu), it converges to a limit gg and it suffices to prove that g=0g=0 a.e. For ψCc(E2,(2)N)\psi\in C^{\infty}_{c}(E_{2},({\mathbb{R}}^{2})^{N}), we have (2)Ngψdμ=limn(2)Nφnψdμ\int_{({\mathbb{R}}^{2})^{N}}g\cdot\psi{\rm d}\mu=\lim_{n}\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi_{n}\cdot\psi{\rm d}\mu. But, recalling (4),

(2)Nφnψdμ=(2)Nφn(x)ψ(x)𝐦(x)dx=(2)Nφn(x)div(𝐦(x)ψ(x))dx.\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi_{n}\cdot\psi{\rm d}\mu=\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi_{n}(x)\cdot\psi(x){\mathbf{m}}(x){\rm d}x=-\int_{({\mathbb{R}}^{2})^{N}}\varphi_{n}(x){\rm div}({\mathbf{m}}(x)\psi(x)){\rm d}x.

Thus by the Cauchy-Schwarz inequality,

|(2)Nφnψdμ|((2)Nφn2dμ)1/2((2)N|div(𝐦(x)ψ(x))|2𝐦(x)dx)1/2,\Big{|}\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi_{n}\cdot\psi{\rm d}\mu\Big{|}\leq\Big{(}\int_{({\mathbb{R}}^{2})^{N}}\varphi_{n}^{2}{\rm d}\mu\Big{)}^{1/2}\Big{(}\int_{({\mathbb{R}}^{2})^{N}}\frac{|{\rm div}({\mathbf{m}}(x)\psi(x))|^{2}}{{\mathbf{m}}(x)}{\rm d}x\Big{)}^{1/2},

which tends to 0 since limnφn=0\lim_{n}\varphi_{n}=0 in L2((2)N,μ)L^{2}(({\mathbb{R}}^{2})^{N},\mu), since ψCc(E2,(2)N)\psi\in C^{\infty}_{c}(E_{2},({\mathbb{R}}^{2})^{N}) and since 𝐦{\mathbf{m}} is smooth and positive on E2E_{2}. Thus (2)Ngψdμ=0\int_{({\mathbb{R}}^{2})^{N}}g\cdot\psi{\rm d}\mu=0 for all ψCc(E2,(2)N)\psi\in C^{\infty}_{c}(E_{2},({\mathbb{R}}^{2})^{N}), so that g=0g=0 a.e.

We can thus consider the extension of X{\mathcal{E}}^{X} to X=Cc(𝒳)¯1X{\mathcal{F}}^{X}=\overline{C_{c}^{\infty}({\mathcal{X}})}^{{\mathcal{E}}^{X}_{1}}, where we have set 1X(φ,φ)=(2)N(φ2+12φ2)dμ{\mathcal{E}}^{X}_{1}(\varphi,\varphi)=\int_{({\mathbb{R}}^{2})^{N}}(\varphi^{2}+\frac{1}{2}||\nabla\varphi||^{2}){\rm d}\mu for φCc(𝒳)\varphi\in C^{\infty}_{c}({\mathcal{X}}).

Next, (X,X)({\mathcal{E}}^{X},{\mathcal{F}}^{X}) is obviously regular with core Cc(𝒳)C^{\infty}_{c}({\mathcal{X}}), see [11, page 6], because Cc(𝒳)C^{\infty}_{c}({\mathcal{X}}) is dense in X{\mathcal{F}}^{X} for the norm associated to 1X{\mathcal{E}}_{1}^{X} by definition of X{\mathcal{F}}^{X} and Cc(𝒳)C^{\infty}_{c}({\mathcal{X}}) is dense, for the uniform norm, in Cc(𝒳)C_{c}({\mathcal{X}}). It is also strongly local, see [11, page 6], i.e. X(φ,ψ)=12(2)Nφψdμ=0{\mathcal{E}}^{X}(\varphi,\psi)=\color[rgb]{0,0,0}\frac{1}{2}\color[rgb]{0,0,0}\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi\cdot\nabla\psi{\rm d}\mu=0 if φ,ψCc(𝒳)\varphi,\psi\in C^{\infty}_{c}({\mathcal{X}}) and if φ\varphi is constant on a neighborhood of Suppψ{\rm Supp}\;\psi.

Then [11, Theorems 7.2.2 page 380 and 4.2.8 page 167] imply the existence and uniqueness of a Hunt process 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳)\mathbb{X}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}_{x}^{X})_{x\in{\mathcal{X}_{\triangle}}}) with values in 𝒳{\mathcal{X}_{\triangle}}, which is μ\mu-symmetric, of which the Dirichlet space is (X,X)({\mathcal{E}}^{X},{\mathcal{F}}^{X}), and such that tXtt\mapsto X_{t} is xX{\mathbb{P}}^{X}_{x}-a.s. continuous on [0,ζ)[0,\zeta) for all x𝒳x\in{\mathcal{X}}, where ζ=inf{t0:Xt=}\zeta=\inf\{t\geq 0:X_{t}=\triangle\}.

Furthermore, since X{\mathcal{E}}^{X} is strongly local, we know from [11, Theorem 4.5.3 page 186] that we can choose 𝕏{\mathbb{X}} (modifying xX{\mathbb{P}}^{X}_{x} only on a properly exceptional set) such that x(ζ<,Xζ=)=1{\mathbb{P}}_{x}(\zeta<\infty,X_{\zeta-}=\triangle)=1 for all x𝒳x\in{\mathcal{X}}. This implies that for all x𝒳x\in{\mathcal{X}}, x{\mathbb{P}}_{x}-a.s., the map tXtt\mapsto X_{t} is continuous from [0,)[0,\infty) to 𝒳{\mathcal{X}}_{\triangle}, endowed with the one-point compactification topology on 𝒳{\mathcal{X}_{\triangle}} recalled in Subsection B.1. Hence 𝕏{\mathbb{X}} is a diffusion.

For (ii), the very same strategy applies. The only difference is the integration by parts to be used for the closability: for φCc1(𝒰)\varphi\in C^{1}_{c}({\mathcal{U}}) and ψCc1(𝕊E2,(2)N)\psi\in C^{1}_{c}({\mathbb{S}}\cap E_{2},({\mathbb{R}}^{2})^{N}), it classically holds that

(1) 𝕊(𝕊φ)ψdβ=𝕊(𝕊φ(u))ψ(u)𝐦(u)σ(du)=𝕊φ(u)div𝕊(𝐦(u)ψ(u))σ(du).\int_{\mathbb{S}}(\nabla_{\mathbb{S}}\varphi)\cdot\psi{\rm d}\beta=\int_{\mathbb{S}}(\nabla_{\mathbb{S}}\varphi(u))\cdot\psi(u){\mathbf{m}}(u)\sigma({\rm d}u)=-\int_{\mathbb{S}}\varphi(u){\rm div}_{\mathbb{S}}({\mathbf{m}}(u)\psi(u))\sigma({\rm d}u).

This can be shown naively using Lemma A.2. ∎

We now make explicit the generators of 𝕏{\mathbb{X}} and 𝕌{\mathbb{U}} when applied to some functions enjoying a few properties. See Subsection B.1 for a precise definition of the generator of a Hunt process. We have to introduce a few notation. For φC((2)N)\varphi\in C^{\infty}(({\mathbb{R}}^{2})^{N}), α(0,1]\alpha\in(0,1] and x(2)Nx\in({\mathbb{R}}^{2})^{N}, we set

(2) αXφ(x)=12Δφ(x)θN1ijNxixjxixj2+α(φ(x))i=12𝐦α(x)div[𝐦α(x)φ(x)],{\mathcal{L}}^{X}_{\alpha}\varphi(x)=\frac{1}{2}\Delta\varphi(x)-\frac{\theta}{N}\sum_{1\leq i\neq j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot(\nabla\varphi(x))^{i}=\frac{1}{2{\mathbf{m}}_{\alpha}(x)}{\rm div}[{\mathbf{m}}_{\alpha}(x)\nabla\varphi(x)],

where

𝐦α(x)=1ijN(xixj2+α)θ/(2N).{\mathbf{m}}_{\alpha}(x)=\prod_{1\leq i\neq j\leq N}(\|x^{i}-x^{j}\|^{2}+\alpha)^{-\theta/(2N)}.

This is in accordance with (4), in the sense that 𝐦0=𝐦{\mathbf{m}}_{0}={\mathbf{m}}. The formula (2) makes sense for xE2x\in E_{2} when α=0\alpha=0 (with 𝐦α{\mathbf{m}}_{\alpha} replaced by 𝐦{\mathbf{m}}) and we recall that for φC((2)N)\varphi\in C^{\infty}(({\mathbb{R}}^{2})^{N}) and xE2x\in E_{2}, Xφ(x){\mathcal{L}}^{X}\varphi(x) was defined in (5) by Xφ(x)=0Xφ(x){\mathcal{L}}^{X}\varphi(x)={\mathcal{L}}^{X}_{0}\varphi(x). We will often use that for all φ,ψC((2)N)\varphi,\psi\in C^{\infty}(({\mathbb{R}}^{2})^{N}), all x(2)Nx\in({\mathbb{R}}^{2})^{N}, all α(0,1]\alpha\in(0,1],

(3) αX(φψ)(x)=φ(x)αXψ(x)+ψ(x)αXφ(x)+φ(x)ψ(x).\displaystyle{\mathcal{L}}_{\alpha}^{X}(\varphi\psi)(x)=\varphi(x){\mathcal{L}}_{\alpha}^{X}\psi(x)+\psi(x){\mathcal{L}}_{\alpha}^{X}\varphi(x)+\nabla\varphi(x)\cdot\nabla\psi(x).

For φC(𝕊)\varphi\in C^{\infty}({\mathbb{S}}), α(0,1]\alpha\in(0,1] and u𝕊u\in{\mathbb{S}}, we set

(4) αUφ(u)=12Δ𝕊φ(u)θN1ijNuiujuiuj2+α(𝕊φ(u))i=12𝐦α(u)div𝕊[𝐦α(u)𝕊φ(u)].{\mathcal{L}}^{U}_{\alpha}\varphi(u)=\frac{1}{2}\Delta_{\mathbb{S}}\varphi(u)-\frac{\theta}{N}\sum_{1\leq i\neq j\leq N}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(\nabla_{\mathbb{S}}\varphi(u))^{i}=\frac{1}{2{\mathbf{m}}_{\alpha}(u)}{\rm div}_{\mathbb{S}}[{\mathbf{m}}_{\alpha}(u)\nabla_{\mathbb{S}}\varphi(u)].

This formula makes sense for u𝕊E2u\in{\mathbb{S}}\cap E_{2} when α=0\alpha=0 (with 𝐦α{\mathbf{m}}_{\alpha} replaced by 𝐦{\mathbf{m}}) and we set, for φC(𝕊)\varphi\in C^{\infty}({\mathbb{S}}) and u𝕊E2u\in{\mathbb{S}}\cap E_{2}, Uφ(u)=0Uφ(u){\mathcal{L}}^{U}\varphi(u)={\mathcal{L}}^{U}_{0}\varphi(u).

Remark 4.3.

(i) Denote by (𝒜X,𝒟AX)({\mathcal{A}}^{X},{\mathcal{D}}_{A^{X}}) the generator of the process 𝕏{\mathbb{X}} of Proposition 4.1-(i). If φCc(𝒳)\varphi\in C^{\infty}_{c}({\mathcal{X}}) satisfies supα(0,1]supx(2)N|αXφ(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}\varphi(x)|<\infty, then φ𝒟AX\varphi\in{\mathcal{D}}_{A^{X}} and 𝒜Xφ=Xφ{\mathcal{A}}^{X}\varphi={\mathcal{L}}^{X}\varphi.

(ii) Denote by (𝒜U,𝒟AU)({\mathcal{A}}^{U},{\mathcal{D}}_{A^{U}}) the generator of the process 𝕌\mathbb{U} of Proposition 4.1-(ii). If φCc(𝒰)\varphi\in C^{\infty}_{c}({\mathcal{U}}) satisfies supα(0,1]supu𝕊|αUφ(u)|<\sup_{\alpha\in(0,1]}\sup_{u\in{\mathbb{S}}}|{\mathcal{L}}^{U}_{\alpha}\varphi(u)|<\infty, then φ𝒟AU\varphi\in{\mathcal{D}}_{A^{U}} and 𝒜Uφ=Uφ{\mathcal{A}}^{U}\varphi={\mathcal{L}}^{U}\varphi.

Proof.

To check (i), it suffices by (1) to verify that (a) φX\varphi\in{\mathcal{F}}^{X}, (b) XφL2(𝒳,μ){\mathcal{L}}^{X}\varphi\in L^{2}({\mathcal{X}},\mu) and (c) for all ψX\psi\in{\mathcal{F}}^{X}, we have X(φ,ψ)=𝒳(Xφ)ψdμ{\mathcal{E}}^{X}(\varphi,\psi)=-\int_{\mathcal{X}}({\mathcal{L}}^{X}\varphi)\psi{\rm d}\mu.

Point (a) is clear, since φCc(𝒳)\varphi\in C^{\infty}_{c}({\mathcal{X}}). Point (b) follows from the facts that μ\mu is Radon on 𝒳{\mathcal{X}}, that φ\varphi is compactly supported in 𝒳{\mathcal{X}} and that XφL((2)N,dx){\mathcal{L}}^{X}\varphi\in L^{\infty}(({\mathbb{R}}^{2})^{N},{\rm d}x), because for all xE2x\in E_{2}, Xφ(x)=limα0αXφ(x){\mathcal{L}}^{X}\varphi(x)=\lim_{\alpha\to 0}{\mathcal{L}}_{\alpha}^{X}\varphi(x). Concerning (c) it suffices, by definition of (X,X)({\mathcal{E}}^{X},{\mathcal{F}}^{X}) and since XφL2(𝒳,μ){\mathcal{L}}^{X}\varphi\in L^{2}({\mathcal{X}},\mu), to show that for all ψCc(𝒳)\psi\in C^{\infty}_{c}({\mathcal{X}}), we have 12(2)Nφψdμ=(2)N(Xφ)ψdμ\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi\cdot\nabla\psi{\rm d}\mu=-\int_{({\mathbb{R}}^{2})^{N}}({\mathcal{L}}^{X}\varphi)\psi{\rm d}\mu. But for α(0,1]\alpha\in(0,1], by a standard integration by parts, since φ,ψ\varphi,\psi and 𝐦α{\mathbf{m}}_{\alpha} are smooth,

12(2)Nφ(x)ψ(x)𝐦α(x)dx=\displaystyle\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\nabla\varphi(x)\cdot\nabla\psi(x){\mathbf{m}}_{\alpha}(x){\rm d}x= 12(2)Ndiv(𝐦α(x)φ(x))ψ(x)dx\displaystyle-\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}{\rm div}({\mathbf{m}}_{\alpha}(x)\nabla\varphi(x))\psi(x){\rm d}x
=\displaystyle= (2)N[αXφ(x)]ψ(x)𝐦α(x)dx.\displaystyle-\int_{({\mathbb{R}}^{2})^{N}}[{\mathcal{L}}^{X}_{\alpha}\varphi(x)]\psi(x){\mathbf{m}}_{\alpha}(x){\rm d}x.

We conclude letting α0\alpha\to 0 by dominated convergence, since 𝐦α𝐦{\mathbf{m}}_{\alpha}\to{\mathbf{m}} and αXφXφ{\mathcal{L}}_{\alpha}^{X}\varphi\to{\mathcal{L}}^{X}\varphi a.e., since by assumption, |φ(x)ψ(x)𝐦α(x)|+|[αXφ(x)]ψ(x)𝐦α(x)|C1I{x𝒦}𝐦(x)|\nabla\varphi(x)\cdot\nabla\psi(x){\mathbf{m}}_{\alpha}(x)|+|[{\mathcal{L}}^{X}_{\alpha}\varphi(x)]\psi(x){\mathbf{m}}_{\alpha}(x)|\leq C\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{x\in{\mathcal{K}}\}}{\mathbf{m}}(x) for some constant CC and for 𝒦={\mathcal{K}}=Supp ψ\psi which is compact in 𝒳{\mathcal{X}}, and since μ(𝒦)=𝒦𝐦(x)dx<\mu({\mathcal{K}})=\int_{\mathcal{K}}{\mathbf{m}}(x){\rm d}x<\infty.

The proof of (ii) is exactly the same, using that if φ,ψC(𝕊)\varphi,\psi\in C^{\infty}({\mathbb{S}}), it holds that

12𝕊𝕊φ𝕊ψ𝐦αdσ=12𝕊div𝕊(𝐦α𝕊φ)ψdσ=𝕊[αUφ]ψ𝐦αdσ,\frac{1}{2}\int_{{\mathbb{S}}}\nabla_{\mathbb{S}}\varphi\cdot\nabla_{\mathbb{S}}\psi\;{\mathbf{m}}_{\alpha}{\rm d}\sigma=-\frac{1}{2}\int_{{\mathbb{S}}}{\rm div}_{\mathbb{S}}({\mathbf{m}}_{\alpha}\nabla_{\mathbb{S}}\varphi)\psi{\rm d}\sigma=-\int_{{\mathbb{S}}}[{\mathcal{L}}^{U}_{\alpha}\varphi]\psi{\mathbf{m}}_{\alpha}{\rm d}\sigma,

which can be shown naively using the projection Φ𝕊\Phi_{\mathbb{S}}, see (3), and Lemma A.2. ∎

We end the section with a quick irreducibility/recurrence/transience study of the spherical process, see Subsection B.1 again for definitions.

Lemma 4.4.

We fix N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta and consider the process 𝕌\mathbb{U} and its Dirichlet space (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) as in Proposition 4.1-(ii).

(i) (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is irreducible and we have the alternative: \bullet either (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is recurrent and in particular it is non-exploding and for all measurable A𝒰A\subset{\mathcal{U}} such that β(A)>0\beta(A)>0, uU(lim supt{UtA})=1{\mathbb{P}}^{U}_{u}(\limsup_{t\to\infty}\{U_{t}\in A\})=1 quasi-everywhere; \bullet or (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is transient and in particular for all compact set 𝒦{\mathcal{K}} of 𝒰{\mathcal{U}}, we have quasi-everywhere uU(lim inft{Ut𝒦})=0{\mathbb{P}}^{U}_{u}(\liminf_{t\to\infty}\{U_{t}\in{\mathcal{K}}\})=0. (ii) If dθ,N(N1)>0d_{\theta,N}(N-1)>0, then (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is recurrent.

In the transient case, one might also prove that uU(lim supt{Ut𝒦})=0{\mathbb{P}}^{U}_{u}(\limsup_{t\to\infty}\{U_{t}\in{\mathcal{K}}\})=0, but this would be useless for our purpose.

Proof.

We start with (i). We first show that in any case, (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is irreducible. By [11, Corollary 4.6.4 page 195] and since U(φ,φ)=12𝕊𝕊φ2𝐦dσ{\mathcal{E}}^{U}(\varphi,\varphi)=\color[rgb]{0,0,0}\frac{1}{2}\color[rgb]{0,0,0}\int_{{\mathbb{S}}}\|\nabla_{\mathbb{S}}\varphi\|^{2}{\mathbf{m}}{\rm d}\sigma with 𝐦{\mathbf{m}} bounded from below by a constant (on 𝕊{\mathbb{S}}), it suffices to prove that the σ\sigma-symmetric Hunt process with regular Dirichlet space (,)({\mathcal{E}},{\mathcal{F}}) on L2(𝒰,σ)L^{2}({\mathcal{U}},\sigma) with core Cc(𝒰)C^{\infty}_{c}({\mathcal{U}}) such that for all φCc(𝒰)\varphi\in C^{\infty}_{c}({\mathcal{U}}), (φ,φ)=12𝕊𝕊φ2dσ{\mathcal{E}}(\varphi,\varphi)=\color[rgb]{0,0,0}\frac{1}{2}\color[rgb]{0,0,0}\int_{{\mathbb{S}}}\|\nabla_{\mathbb{S}}\varphi\|^{2}{\rm d}\sigma is irreducible. But this Hunt process is nothing but a 𝕊{\mathbb{S}}-valued Brownian motion. This Brownian motion is a priori killed when it gets out of 𝒰{\mathcal{U}}, but this does a.s. never occur since such a Brownian motion never has two (bi-dimensional) coordinates equal. This 𝕊{\mathbb{S}}-valued Brownian motion is of course irreducible. We conclude from [11, Lemma 1.6.4 page 55] that (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is either recurrent or transient.

\bullet When (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is recurrent, [11, Theorem 4.7.1-(iii) page 202] gives us the result.

\bullet When (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is transient, we fix a compact set 𝒦{\mathcal{K}} of 𝒰{\mathcal{U}} and we know from Lemma A.3 that β(𝒦)<\beta({\mathcal{K}})<\infty, so that by definition of transience, for β\beta-a.e u𝒰u\in{\mathcal{U}}, 𝔼uU[01I𝒦(Us)ds]<{\mathbb{E}}^{U}_{u}[\int_{0}^{\infty}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\mathcal{K}}(U_{s}){\rm d}s]<\infty. Setting τ𝒦c=inf{t0:Ut𝒦}\tau_{{\mathcal{K}}^{c}}=\inf\{t\geq 0:U_{t}\notin{\mathcal{K}}\}, we get in particular that for β\beta-a.e u𝒰u\in{\mathcal{U}}, uU(τ𝒦c<)=1{\mathbb{P}}^{U}_{u}(\tau_{{\mathcal{K}}^{c}}<\infty)=1. But, by [11, (4.1.9) page 155], uuU(τ𝒦c<)u\mapsto{\mathbb{P}}^{U}_{u}(\tau_{{\mathcal{K}}^{c}}<\infty) is finely continuous. Using [11, Lemma 4.1.5 page 155], we deduce that uU(τ𝒦c<)=1{\mathbb{P}}^{U}_{u}(\tau_{{\mathcal{K}}^{c}}<\infty)=1 quasi-everywhere. The Markov property allows us to conclude.

Concerning (ii), we recall from Proposition A.3 that β(𝕊)<\beta({\mathbb{S}})<\infty, because dθ,N(N1)>0d_{\theta,N}(N-1)>0 implies that k0Nk_{0}\geq N, see Lemma 1.1. Moreover, k0Nk_{0}\geq N implies that Ek0EN𝕊E_{k_{0}}\supset E_{N}\supset{\mathbb{S}}, whence 𝒰=Ek0𝕊=𝕊{\mathcal{U}}=E_{k_{0}}\cap{\mathbb{S}}={\mathbb{S}} is compact: the process cannot explode, i.e. ξ=\xi=\infty. Consequently, (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is recurrent, since φ1\varphi\equiv 1 belongs to L1(𝒰,β)L^{1}({\mathcal{U}},\beta) and since 𝔼uU[0φ(Us)ds]=𝔼uU[ξ]={\mathbb{E}}_{u}^{U}[\int_{0}^{\infty}\varphi(U_{s}){\rm d}s]={\mathbb{E}}_{u}^{U}[\xi]=\infty. Indeed, as recalled Subsection B.1, if (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) was transient, we would have 𝔼uU[0φ(Us)ds]<{\mathbb{E}}_{u}^{U}[\int_{0}^{\infty}\varphi(U_{s}){\rm d}s]<\infty for all φL1(𝒰,β)\varphi\in L^{1}({\mathcal{U}},\beta), with the convention that φ()=0\varphi(\triangle)=0. ∎

5. Decomposition

The goal of this section is to prove the following decomposition of the Keller-Segel particle system defined in Proposition 4.1-(i). This decomposition is noticeable and crucial for our purpose.

Proposition 5.1.

We fix N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta, and we recall that k0=2N/θk_{0}=\lceil 2N/\theta\rceil, that 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} and that 𝒰=𝕊Ek0{\mathcal{U}}={\mathbb{S}}\cap E_{k_{0}}.

For xENx\in E_{N}, we set r=R[[1,N]](x)>0r=R_{[\![1,N]\!]}(x)>0, z=S[[1,N]](x)2z=S_{[\![1,N]\!]}(x)\in{\mathbb{R}}^{2} and u=(xγ(z))/r𝕊u=(x-\gamma(z))/\sqrt{r}\in{\mathbb{S}} and we consider three independent processes: \bullet (Mt)t0(M_{t})_{t\geq 0}, a 22-dimensional Brownian motion with diffusion constant N1/2N^{-1/2} starting from zz, \bullet (Dt)t0(D_{t})_{t\geq 0} a squared Bessel process with dimension dθ,N(N)d_{\theta,N}(N) starting from rr and killed when it gets out of (0,)(0,\infty), with life-time τD=inf{t0:Dt=}\tau_{D}=\inf\{t\geq 0:D_{t}=\triangle\}, \bullet (Ut)t0(U_{t})_{t\geq 0}, a QSKS(θ,N)QSKS(\theta,N) -process starting from uu, with life-time ξ=inf{t0:Ut=}\xi=\inf\{t\geq 0:U_{t}=\triangle\}. We introduce At=0tτDDs1dsA_{t}=\int_{0}^{t\land\tau_{D}}D_{s}^{-1}{\rm d}s, and its generalized inverse ρt=inf{s>0:As>t}\rho_{t}=\inf\{s>0:A_{s}>t\}. We define Yt=Ψ(Mt,Dt,UAt)Y_{t}=\Psi(M_{t},D_{t},U_{A_{t}}), where we recall from (2) that Ψ(z,r,u)=γ(z)+ruEN\Psi(z,r,u)=\gamma(z)+\sqrt{r}u\in E_{N} when (z,r,u)2×(0,)×𝕊(z,r,u)\in{\mathbb{R}}^{2}\times(0,\infty)\times{\mathbb{S}} and where we set Ψ(z,r,u)=\Psi(z,r,u)=\triangle when r=r=\triangle or u=u=\triangle. Observe that the life-time of YY equals ζ=ρξτD\zeta^{\prime}=\rho_{\xi}\land\tau_{D}. Consider also a QKS(θ,N)QKS(\theta,N)-process 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳){\mathbb{X}}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}_{x}^{X})_{x\in{\mathcal{X}_{\triangle}}}), with life-time ζ\zeta, and 𝕏=(ΩX,X,(Xt)t0,(xX)x(𝒳EN){}){\mathbb{X}^{*}}=(\Omega^{X},{\mathcal{M}}^{X},(X^{*}_{t})_{t\geq 0},({\mathbb{P}}_{x}^{X})_{x\in({\mathcal{X}}\cap E_{N})\cup\{\triangle\}}), where Xt=Xt1I{t<τ}+1I{tτ}X^{*}_{t}=X_{t}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{t<\tau\}}+\triangle\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{t\geq\tau\}} and where τ=inf{t0:R[[1,N]](Xt)(0,)}\tau=\inf\{t\geq 0:R_{[\![1,N]\!]}(X_{t})\notin(0,\infty)\}. In other words, 𝕏{\mathbb{X}^{*}} is the version of 𝕏{\mathbb{X}} killed when it gets out of ENE_{N}. The life-time of 𝕏{\mathbb{X}^{*}} is τ\tau.

The law of (Yt)t0(Y_{t})_{t\geq 0} is the same as that of (Xt)t0(X^{*}_{t})_{t\geq 0} under xX{\mathbb{P}}_{x}^{X}, quasi-everywhere in 𝒳EN{\mathcal{X}}\cap E_{N}.

We take the convention that R[[1,N]]()=0R_{[\![1,N]\!]}(\triangle)=0, so that τ[0,ζ]\tau\in[0,\zeta]. Since R[[1,N]](Yt)=DtR_{[\![1,N]\!]}(Y_{t})=D_{t} and S[[1,N]](Yt)=MtS_{[\![1,N]\!]}(Y_{t})=M_{t} for all t[0,ζ)t\in[0,\zeta^{\prime}), Proposition 5.1 in particular implies that (R[[1,N]](Xt))t0(R_{[\![1,N]\!]}(X_{t}))_{t\geq 0} and (S[[1,N]](Xt))t0(S_{[\![1,N]\!]}(X_{t}))_{t\geq 0} are some independent squared Bessel process and Brownian motion until the first time (R[[1,N]](Xt))t0(R_{[\![1,N]\!]}(X_{t}))_{t\geq 0} vanishes. This actually holds true until explosion, as shown in Lemma 5.2 below. The quasi-everywhere notion refers to the Hunt process 𝕏\mathbb{X}. Observe that when θ2\theta\geq 2, we have k0Nk_{0}\leq N, so that 𝒳EN=𝒳{\mathcal{X}}\cap E_{N}={\mathcal{X}} and 𝕏=𝕏{\mathbb{X}}={\mathbb{X}}^{*}.

Proof.

We slice the proof in several steps. The two first steps are more or less classical, even if we give all the details: we determine the Dirichlet spaces of the three processes (Mt)t0(M_{t})_{t\geq 0}, (Dt)t0(D_{t})_{t\geq 0} and (Ut)t0(U_{t})_{t\geq 0} involved in the construction of (Yt)t0(Y_{t})_{t\geq 0}; then we compute the Dirichlet space of (Dρt)t0(D_{\rho_{t}})_{t\geq 0}; we next identify the Dirichlet space of (Dρt,Ut)t0(D_{\rho_{t}},U_{t})_{t\geq 0}, which allows us to find the one of (Dt,UAt)t0(D_{t},U_{A_{t}})_{t\geq 0} by a second time-change; by concatenation, we deduce the Dirichlet space of (Mt,Dt,UAt)t0(M_{t},D_{t},U_{A_{t}})_{t\geq 0}. The main computations are handled in Steps 3 and 4, where we find the Dirichlet space of (Yt)t0(Y_{t})_{t\geq 0}, which allows us to conclude in Step 5 by uniqueness.

Step 1. First, take 𝕌=(ΩU,U,(Ut)t0,(uU)u𝒰)\mathbb{U}=(\Omega^{U},{\mathcal{M}}^{U},(U_{t})_{t\geq 0},({\mathbb{P}}^{U}_{u})_{u\in{\mathcal{U}_{\triangle}}}) as in Proposition 4.1-(ii).

Second, consider a 22-dimensional Brownian motion 𝕄=(ΩM,M,(Mt)t0,(zM)z2)\mathbb{M}=(\Omega^{M},{\mathcal{M}}^{M},(M_{t})_{t\geq 0},({\mathbb{P}}^{M}_{z})_{z\in\mathbb{R}^{2}}) with diffusion constant N1/2N^{-1/2}. We know from [11, Example 4.2.1 page 167] that 𝕄\mathbb{M} is a dz{\rm d}z-symmetric (here dz{\rm d}z is the Lebesgue measure on 2{\mathbb{R}}^{2}) diffusion with regular Dirichlet space (M,M)({\mathcal{E}}^{M},{\mathcal{F}}^{M}) on L2(2,dz)L^{2}({\mathbb{R}}^{2},{\rm d}z) with core Cc(2)C_{c}^{\infty}({\mathbb{R}}^{2}) and for all φCc(2)\varphi\in C_{c}^{\infty}({\mathbb{R}}^{2}),

(1) M(φ,φ)=12N2zφ(z)2dz.\mathcal{E}^{M}(\varphi,\varphi)=\frac{1}{2N}\int_{{\mathbb{R}}^{2}}\|\nabla_{z}\varphi(z)\|^{2}{\rm d}z.

Finally, let 𝔻=(ΩD,D,(Dt)t0,(rD)r+{})\mathbb{D}=(\Omega^{D},{\mathcal{M}}^{D},(D_{t})_{t\geq 0},({\mathbb{P}}^{D}_{r})_{r\in{\mathbb{R}}_{+}^{*}\cup\{\triangle\}}) be a squared Bessel process of dimension dθ,N(N)d_{\theta,N}(N) killed when it gets out of +=(0,){\mathbb{R}}_{+}^{*}=(0,\infty) and set ν=dθ,N(N)/21\nu=d_{\theta,N}(N)/2-1, see Revuz-Yor [21, page 443]. Fukushima [10, Theorem 3.3] tells us that 𝔻\mathbb{D} is a rνdrr^{\nu}{\rm d}r-symmetric diffusion (here dr{\rm d}r is the Lebesgue measure on +{\mathbb{R}}_{+}^{*}) with regular Dirichlet space (D,D)({\mathcal{E}}^{D},{\mathcal{F}}^{D}) on L2(+,rνdr)L^{2}({\mathbb{R}}_{+},r^{\nu}{\rm d}r) with core Cc(+)C_{c}^{\infty}({\mathbb{R}}_{+}^{*}) where for all φCc(+)\varphi\in C_{c}^{\infty}({\mathbb{R}}_{+}^{*}),

(2) D(φ,φ)=2+|φ(r)|2rν+1dr.\mathcal{E}^{D}(\varphi,\varphi)=2\int_{{\mathbb{R}}_{+}}|\varphi^{\prime}(r)|^{2}r^{\nu+1}{\rm d}r.

Together with [10, Theorem 3.3], this uses that the scale function and the speed measure of (Dt)t0(D_{t})_{t\geq 0} are respectively rrνr\mapsto r^{-\nu} and [rν/(2ν)]dr-[r^{\nu}/(2\nu)]{\rm d}r. Actually, we don’t take the speed measure as reference measure but rνdrr^{\nu}{\rm d}r which is the same up to a constant.

Step 2. We apply Lemma B.3 to 𝔻\mathbb{D} with g(r)=1/rg(r)=1/r, i.e. with At=0tDs1ds=0tτDDs1dsA_{t}=\int_{0}^{t}D_{s}^{-1}{\rm d}s=\int_{0}^{t\land\tau_{D}}D_{s}^{-1}{\rm d}s thanks to the convention 1=0\triangle^{-1}=0 and recall that ρ\rho is its generalized inverse: we find that setting Dρt=Dρt1I{ρt<}+1I{ρt=}D_{\rho_{t}}=D_{\rho_{t}}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\rho_{t}<\infty\}}+\triangle\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\rho_{t}=\infty\}},

𝔻ρ=(ΩD,D,(Dρt)t0,(rD)r+)\mathbb{D}_{\rho}=(\Omega^{D},{\mathcal{M}}^{D},(D_{\rho_{t}})_{t\geq 0},({\mathbb{P}}^{D}_{r})_{r\in{\mathbb{R}}^{*}_{+}})

is a rν1drr^{\nu-1}{\rm d}r-symmetric (+{})({\mathbb{R}}^{*}_{+}\cup\{\triangle\})-valued diffusion with regular Dirichlet space (Dρ,Dρ)({\mathcal{E}}^{D_{\rho}},{\mathcal{F}}^{D_{\rho}}) on L2(+,rν1dr)L^{2}({\mathbb{R}}_{+},r^{\nu-1}{\rm d}r) with core Cc(+)C_{c}^{\infty}({\mathbb{R}}^{*}_{+}) such that for all φCc(+)\varphi\in C_{c}^{\infty}({\mathbb{R}}^{*}_{+}),

(3) Dρ(φ,φ)=D(φ,φ)=2+|φ(r)|2rν+1dr=2+|rφ(r)|2rν1dr.\mathcal{E}^{D_{\rho}}(\varphi,\varphi)=\mathcal{E}^{D}(\varphi,\varphi)=2\int_{{\mathbb{R}}_{+}}|\varphi^{\prime}(r)|^{2}r^{\nu+1}{\rm d}r=2\int_{{\mathbb{R}}_{+}}|r\varphi^{\prime}(r)|^{2}r^{\nu-1}{\rm d}r.

We use Lemma B.5 and the notation therein: recalling that (D,U)=σ((Dρt,Ut):t0){\mathcal{M}}^{(D,U)}=\sigma((D_{\rho_{t}},U_{t}):t\geq 0), with the convention that (r,)=(,u)=(,)=(r,\triangle)=(\triangle,u)=(\triangle,\triangle)=\triangle, and that (r,u)(D,U)=rDuU{\mathbb{P}}^{(D,U)}_{(r,u)}={\mathbb{P}}^{D}_{r}\otimes{\mathbb{P}}^{U}_{u} if (r,u)+×𝒰(r,u)\in\mathbb{R}^{*}_{+}\times{\mathcal{U}} and (D,U)=DU{\mathbb{P}}^{(D,U)}_{\triangle}={\mathbb{P}}^{D}_{\triangle}\otimes{\mathbb{P}}^{U}_{\triangle}, it holds that

(𝔻ρ,𝕌)=(ΩD×ΩU,(D,U),(Dρt,Ut)t0,((r,u)(D,U))(r,u)(+×𝒰){})(\mathbb{D_{\rho}},\mathbb{U})=\Big{(}\Omega^{D}\times\Omega^{U},{\mathcal{M}}^{(D,U)},(D_{\rho_{t}},U_{t})_{t\geq 0},({\mathbb{P}}^{(D,U)}_{(r,u)})_{(r,u)\in(\mathbb{R}^{*}_{+}\times{\mathcal{U}})\cup\{\triangle\}}\Big{)}

is a rν1drβ(du)r^{\nu-1}{\rm d}r\beta({\rm d}u)-symmetric (+×𝒰){}(\mathbb{R}^{*}_{+}\times{\mathcal{U}})\cup\{\triangle\}-valued diffusion with regular Dirichlet space given by ((Dρ,U),(Dρ,U))(\mathcal{E}^{(D_{\rho},U)},\mathcal{F}^{(D_{\rho},U)}) on L2(+×𝕊,rν1drβ(du))L^{2}({\mathbb{R}}_{+}\times{\mathbb{S}},r^{\nu-1}{\rm d}r\beta({\rm d}u)) with core Cc(+×𝒰)C_{c}^{\infty}({\mathbb{R}}_{+}^{*}\times{\mathcal{U}}), and for all φCc(+×𝒰)\varphi\in C_{c}^{\infty}({\mathbb{R}}_{+}^{*}\times{\mathcal{U}}),

(Dρ,U)(φ,φ)=+U(φ(r,),φ(r,))rν1dr+𝕊Dρ(φ(,u),φ(,u))β(du).\mathcal{E}^{(D_{\rho},U)}(\varphi,\varphi)=\int_{{\mathbb{R}}_{+}}\mathcal{E}^{U}(\varphi(r,\cdot),\varphi(r,\cdot))r^{\nu-1}{\rm d}r+\int_{{\mathbb{S}}}\mathcal{E}^{D_{\rho}}(\varphi(\cdot,u),\varphi(\cdot,u))\beta({\rm d}u).

We now apply Lemma B.3 to (𝔻ρ,𝕌)(\mathbb{D_{\rho}},\mathbb{U}) with g(r,u)=rg(r,u)=r for all r+r\in{\mathbb{R}}_{+}^{*} and all u𝒰u\in{\mathcal{U}}. We consider the time-change αt=0tg(Dρs,Us)ds\alpha_{t}=\int_{0}^{t}g(D_{\rho_{s}},U_{s}){\rm d}s, with the convention that g(r,u)=0g(r,u)=0 as soon as (r,u)=(r,u)=\triangle. We also set Bt=inf{s>0:αs>t}B_{t}=\inf\{s>0:\alpha_{s}>t\}. As we will see in a few lines, it holds that

(4) (DρBt,UBt)=(Dt,UAt)for all t0.(D_{\rho_{B_{t}}},U_{B_{t}})=(D_{t},U_{A_{t}})\qquad\hbox{for all $t\geq 0$}.

Hence Lemma B.3 tells us that

(𝔻,𝕌A)=(ΩD×ΩU,(D,U),(Dt,UAt)t0,((r,u)(D,U))(r,u)(+×𝒰){})(\mathbb{D},\mathbb{U}_{A})=\Big{(}\Omega^{D}\times\Omega^{U},{\mathcal{M}}^{(D,U)},(D_{t},U_{A_{t}})_{t\geq 0},({\mathbb{P}}^{(D,U)}_{(r,u)})_{(r,u)\in(\mathbb{R}^{*}_{+}\times{\mathcal{U}})\cup\{\triangle\}}\Big{)}

is a rνdrβ(du)r^{\nu}{\rm d}r\beta({\rm d}u)-symmetric (+×𝒰){}(\mathbb{R}^{*}_{+}\times{\mathcal{U}})\cup\{\triangle\}-valued diffusion with Dirichlet space ((D,UA),(D,UA))({\mathcal{E}}^{(D,U_{A})},{\mathcal{F}}^{(D,U_{A})}) on L2(+×𝕊,rνdrβ(du))L^{2}({\mathbb{R}}_{+}\times{\mathbb{S}},r^{\nu}{\rm d}r\beta({\rm d}u)), regular with core Cc(+×𝒰)C_{c}^{\infty}({\mathbb{R}}^{*}_{+}\times{\mathcal{U}}) and for all φCc(+×𝒰)\varphi\in C_{c}^{\infty}({\mathbb{R}}^{*}_{+}\times{\mathcal{U}}),

(5) (D,UA)(φ,φ)=(Dρ,U)(φ,φ)=+U(φ(r,),φ(r,))rν1dr+𝕊Dρ(φ(,u),φ(,u))β(du).\displaystyle{\mathcal{E}}^{(D,U_{A})}(\varphi,\varphi)=\mathcal{E}^{(D_{\rho},U)}(\varphi,\varphi)=\int_{{\mathbb{R}}_{+}}\!\!\mathcal{E}^{U}(\varphi(r,\cdot),\varphi(r,\cdot))r^{\nu-1}{\rm d}r+\int_{{\mathbb{S}}}\!\!\mathcal{E}^{D_{\rho}}(\varphi(\cdot,u),\varphi(\cdot,u))\beta({\rm d}u).

We now check the claim (4). Recall that DD explodes at time τD\tau_{D}, that At=0tτDDs1dsA_{t}=\int_{0}^{t\land\tau_{D}}D_{s}^{-1}{\rm d}s and that ρ\rho is the generalized inverse of AA. Hence (ρt)t[0,AτD)(\rho_{t})_{t\in[0,A_{\tau_{D}})} is the true inverse of (At)t[0,τD)(A_{t})_{t\in[0,\tau_{D})} and we have ρt=Dρt\rho_{t}^{\prime}=D_{\rho_{t}}, whence ρt=0tDρsds\rho_{t}=\int_{0}^{t}D_{\rho_{s}}{\rm d}s for t[0,AτD)t\in[0,A_{\tau_{D}}). We also have ρt=\rho_{t}=\infty for tAτDt\geq A_{\tau_{D}}. Next, αt=0tDρsds=ρt\alpha_{t}=\int_{0}^{t}D_{\rho_{s}}{\rm d}s=\rho_{t} for t[0,AτDξ)t\in[0,A_{\tau_{D}}\land\xi), because g(Dρs,Us)=Dρsg(D_{\rho_{s}},U_{s})=D_{\rho_{s}} if (Dρs,Us)(D_{\rho_{s}},U_{s})\neq\triangle, i.e. if s<AτDξs<A_{\tau_{D}}\land\xi. Hence BB, the generalized inverse of α\alpha, equals AA during [0,τDρξ)[0,\tau_{D}\land\rho_{\xi}), thus in particular ρBt=t\rho_{B_{t}}=t for t[0,AτDξ)t\in[0,A_{\tau_{D}}\land\xi). As conclusion, (4) holds true for t[0,AτDξ)t\in[0,A_{\tau_{D}}\land\xi). If now tτDρξt\geq\tau_{D}\land\rho_{\xi}, then Bt=B_{t}=\infty, because BB is the generalized inverse of α\alpha and because for all t0t\geq 0,

αtαAτDξ=ρAτDξ=τDρξ.\alpha_{t}\leq\alpha_{A_{\tau_{D}}\land\xi}=\rho_{A_{\tau_{D}}\land\xi}=\tau_{D}\land\rho_{\xi}.

Hence, still if tτDρξt\geq\tau_{D}\land\rho_{\xi}, we have (DρBt,UBt)=(D_{\rho_{B_{t}}},U_{B_{t}})=\triangle, while (Dt,UAt)=(D_{t},U_{A_{t}})=\triangle because either tτDt\geq\tau_{D} and thus Dt=D_{t}=\triangle or tρξt\geq\rho_{\xi} and thus AtξA_{t}\geq\xi so that UAt=U_{A_{t}}=\triangle. We have proved (4).

We finally conclude, thanks to Lemma B.5 again, setting (M,D,U)=σ((Mt,Dt,UAt):t0){\mathcal{M}}^{(M,D,U)}=\sigma((M_{t},D_{t},U_{A_{t}}):t\geq 0) with the convention that (z,)=(z,\triangle)=\triangle and setting (z,r,u)(M,D,U)=zM(r,u)(D,U){\mathbb{P}}_{(z,r,u)}^{(M,D,U)}={\mathbb{P}}^{M}_{z}\otimes{\mathbb{P}}^{(D,U)}_{(r,u)} in the case where (z,r,u)2×+×𝒰{\color[rgb]{0,0,0}(z,r,u)\color[rgb]{0,0,0}\in\mathbb{R}^{2}\times\mathbb{R}^{*}_{+}\times{\mathcal{U}}} and (M,D,U)=M(D,U){\mathbb{P}}_{\triangle}^{(M,D,U)}={\mathbb{P}}^{M}_{\triangle}\otimes{\mathbb{P}}^{(D,U)}_{\triangle}, that

(𝕄,𝔻,𝕌𝔸)=(ΩM×ΩD×ΩU,(M,D,U),(Mt,Dt,UAt)t0,((z,r,u)(M,D,U))(z,r,u)(2×+×𝒰){})\displaystyle(\mathbb{M},\mathbb{D},\mathbb{U_{A}})=\Big{(}\Omega^{M}\times\Omega^{D}\times\Omega^{U},{\mathcal{M}}^{(M,D,U)},(M_{t},D_{t},U_{A_{t}})_{t\geq 0},({\mathbb{P}}_{(z,r,u)}^{(M,D,U)})_{(z,r,u)\in(\mathbb{R}^{2}\times\mathbb{R}^{*}_{+}\times{\mathcal{U}})\cup\{\triangle\}}\Big{)}

is a dzrνdrβ(du){\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u)-symmetric (2×+×𝒰){}(\mathbb{R}^{2}\times\mathbb{R}^{*}_{+}\times{\mathcal{U}})\cup\{\triangle\}-valued diffusion with regular Dirichlet space ((M,D,UA),(M,D,UA))({\mathcal{E}}^{(M,D,U_{A})},{\mathcal{F}}^{(M,D,U_{A})}) on L2(2×+×𝕊,dzrνdrβ(du))L^{2}({\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}},{\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u)), with core Cc(2×+×𝒰)C_{c}^{\infty}({\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathcal{U}}). Moreover, for all φCc(2×+×𝒰)\varphi\in C_{c}^{\infty}({\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathcal{U}}),

(M,D,UA)(φ,\displaystyle{\mathcal{E}}^{(M,D,U_{A})}(\varphi, φ)=+×𝕊M(φ(,r,u),φ(,r,u))rνdrβ(du)+2(D,UA)(φ(z,,),φ(z,,))dz\displaystyle\varphi)=\int_{{\mathbb{R}}_{+}\times{\mathbb{S}}}\!\!{\mathcal{E}}^{M}(\varphi(\cdot,r,u),\varphi(\cdot,r,u))r^{\nu}{\rm d}r\beta({\rm d}u)+\int_{{\mathbb{R}}^{2}}\!\!\mathcal{E}^{(D,U_{A})}(\varphi(z,\cdot,\cdot),\varphi(z,\cdot,\cdot)){\rm d}z
=\displaystyle= +×𝕊M(φ(,r,u),φ(,r,u))rνdrβ(du)+2×𝕊Dρ(φ(z,,u),φ(z,,u))dzβ(du)\displaystyle\int_{{\mathbb{R}}_{+}\times{\mathbb{S}}}\!\!{\mathcal{E}}^{M}(\varphi(\cdot,r,u),\varphi(\cdot,r,u))r^{\nu}{\rm d}r\beta({\rm d}u)+\int_{{\mathbb{R}}^{2}\times{\mathbb{S}}}\!\!\mathcal{E}^{D_{\rho}}(\varphi(z,\cdot,u),\varphi(z,\cdot,u)){\rm d}z\beta({\rm d}u)
+2×+U(φ(z,r,),φ(z,r,))dzrν1dr\displaystyle\hskip 19.91684pt+\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}}\mathcal{E}^{U}(\varphi(z,r,\cdot),\varphi(z,r,\cdot)){\rm d}zr^{\nu-1}{\rm d}r
(6) =\displaystyle= 2×+×𝕊[12Nzφ(z,r,u)2+2r|rφ(z,r,u)|2+12r𝕊φ(z,r,u)2]dzrνdrβ(du).\displaystyle\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}\Big{[}\frac{1}{2N}||\nabla_{z}\varphi(z,r,u)||^{2}+2r|\partial_{r}\varphi(z,r,u)|^{2}+\frac{1}{2r}||\nabla_{\mathbb{S}}\varphi(z,r,u)||^{2}\Big{]}{\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u).

For the second line, we used (5). For the last line, we used (1), (3) and the expression of U{\mathcal{E}}^{U}, see Proposition 4.1-(ii).

Step 3. We recall that Yt=Ψ(Mt,Dt,UAt)Y_{t}=\Psi(M_{t},D_{t},U_{A_{t}}), where Ψ(z,r,u)=γ(z)+ru\Psi(z,r,u)=\gamma(z)+\sqrt{r}u for (z,r,u)2×+×𝒰(z,r,u)\in{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}^{*}\times{\mathcal{U}} and Ψ(z,r,u)=\Psi(z,r,u)=\triangle for (z,r,u)=(z,r,u)=\triangle. One easily checks that Ψ\Psi is a bijection from (2×+×𝒰){}({\mathbb{R}}^{2}\times{\mathbb{R}}_{+}^{*}\times{\mathcal{U}})\cup\{\triangle\} to (𝒳EN){}({\mathcal{X}}\cap E_{N})\cup\{\triangle\}, recall that 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} and 𝒰=Ek0𝕊{\mathcal{U}}=E_{k_{0}}\cap{\mathbb{S}}.

We now study

𝕐=(ΩY,Y,(Yt)t0,(yY)y(𝒳EN){}),\mathbb{Y}=(\Omega^{Y},{\mathcal{M}}^{Y},(Y_{t})_{t\geq 0},({\mathbb{P}}^{Y}_{y})_{y\in({\mathcal{X}}\cap E_{N})\cup\{\triangle\}}),

where ΩY=ΩM×ΩD×ΩU\Omega^{Y}=\Omega^{M}\times\Omega^{D}\times\Omega^{U}, Y=(M,D,U){\mathcal{M}}^{Y}={\mathcal{M}}^{(M,D,U)} and yY=(z,r,u)(M,D,U){\mathbb{P}}^{Y}_{y}={\mathbb{P}}_{(z,r,u)}^{(M,D,U)} for (z,r,u)=Ψ1(y)(z,r,u)=\Psi^{-1}(y).

First, 𝕐\mathbb{Y} is a (𝒳EN){}({\mathcal{X}}\cap E_{N})\cup\{\triangle\}-valued diffusion, because the bijection Ψ\Psi from (2×+×𝒰){}({\mathbb{R}}^{2}\times{\mathbb{R}}_{+}^{*}\times{\mathcal{U}})\cup\{\triangle\} to (𝒳EN){}({\mathcal{X}}\cap E_{N})\cup\{\triangle\} is continuous, both sets being endowed with the one-point compactification topology, see Subsection B.1.

Next, we prove that 𝕐\mathbb{Y} is μ\mu-symmetric: if φ,ψ\varphi,\psi are nonnegative measurable functions on 𝒳EN{\mathcal{X}}\cap E_{N} and t0t\geq 0, we have, thanks to Lemma A.2 (recall that ν=dθ,N(N)/21\nu=d_{\theta,N}(N)/2-1),

(2)N[PtYφ(y)]ψ(y)μ(dy)=122×+×𝕊[(PtYφ)(Ψ(z,r,u))]ψ(Ψ(z,r,u))rνdzdrβ(du).\int_{({\mathbb{R}}^{2})^{N}}[P^{Y}_{t}\varphi(y)]\psi(y)\mu({\rm d}y)=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}[(P^{Y}_{t}\varphi)(\Psi(z,r,u))]\psi(\Psi(z,r,u))r^{\nu}{\rm d}z{\rm d}r\beta({\rm d}u).

But (PtYφ)(Ψ(z,r,u))=𝔼(z,r,u)[φ(Ψ(Mt,Dt,UAt))]=Pt(M,D,UA)(φΨ)(z,r,u)(P^{Y}_{t}\varphi)(\Psi(z,r,u))=\mathbb{E}_{(z,r,u)}[\varphi(\Psi(M_{t},D_{t},U_{A_{t}}))]=P^{(M,D,U_{A})}_{t}(\varphi\circ\Psi)(z,r,u), so that

(2)N[PtYφ(y)]ψ(y)μ(dy)=\displaystyle\int_{({\mathbb{R}}^{2})^{N}}[P^{Y}_{t}\varphi(y)]\psi(y)\mu({\rm d}y)= 122×+×𝕊[Pt(M,D,UA)(φΨ)(z,r,u)][(ψΨ)(z,r,u)]rνdzdrβ(du).\displaystyle\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}[P^{(M,D,U_{A})}_{t}(\varphi\circ\Psi)(z,r,u)][(\psi\circ\Psi)(z,r,u)]r^{\nu}{\rm d}z{\rm d}r\beta({\rm d}u).

Using that (𝕄,𝔻,𝕌𝔸)(\mathbb{M},\mathbb{D},\mathbb{U_{A}}) is dzrνdrβ(du){\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u)-symmetric and then the same computation in reverse order, one concludes that (2)N[PtYφ]ψdμ=(2)Nφ[PtYψ]dμ\int_{({\mathbb{R}}^{2})^{N}}[P^{Y}_{t}\varphi]\psi{\rm d}\mu=\int_{({\mathbb{R}}^{2})^{N}}\varphi[P^{Y}_{t}\psi]{\rm d}\mu as desired.

Thus 𝕐\mathbb{Y} has a Dirichlet space (Y,Y)({\mathcal{E}}^{Y},{\mathcal{F}}^{Y}) on L2((2)N,μ)L^{2}(({\mathbb{R}}^{2})^{N},\mu) that we now determine. For φL2((2)N,μ)\varphi\in L^{2}(({\mathbb{R}}^{2})^{N},\mu), using as above Lemma A.2 and that (PtYφ)(Ψ(z,r,u))=Pt(M,D,UA)(φΨ)(z,r,u)(P^{Y}_{t}\varphi)(\Psi(z,r,u))=P^{(M,D,U_{A})}_{t}(\varphi\circ\Psi)(z,r,u),

1t(2)N(PtYφφ)φdμ\displaystyle\frac{1}{t}\int_{({\mathbb{R}}^{2})^{N}}(P^{Y}_{t}\varphi-\varphi)\varphi{\rm d}\mu
=\displaystyle= 12t2×+×𝕊[Pt(M,D,UA)(φΨ)(z,r,u)(φΨ)(z,r,u)][φΨ(z,r,u)]rνdzdrβ(du).\displaystyle\frac{1}{2t}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathbb{S}}}[P^{(M,D,U_{A})}_{t}(\varphi\circ\Psi)(z,r,u)-(\varphi\circ\Psi)(z,r,u)][\varphi\circ\Psi(z,r,u)]r^{\nu}{\rm d}z{\rm d}r\beta({\rm d}u).

Since Ψ\Psi is bijective, we deduce, see [11, Lemma 1.3.4 page 23], that

(7) Y={φL2((2)N,μ):φΨ(M,D,UA)}\displaystyle{\mathcal{F}}^{Y}=\Big{\{}\varphi\in L^{2}(({\mathbb{R}}^{2})^{N},\mu):\varphi\circ\Psi\in{\mathcal{F}}^{(M,D,U_{A})}\Big{\}}
(8) and for φY,Y(φ,φ)=12(M,D,UA)(φΨ,φΨ).\displaystyle\hbox{and for $\varphi\in{\mathcal{F}}^{Y}$,}\quad{\mathcal{E}}^{Y}(\varphi,\varphi)=\frac{1}{2}{\mathcal{E}}^{(M,D,U_{A})}(\varphi\circ\Psi,\varphi\circ\Psi).

Step 4. We now compute Y(φ,φ){\mathcal{E}}^{Y}(\varphi,\varphi) for φCc(𝒳EN)\varphi\in C^{\infty}_{c}({\mathcal{X}}\cap E_{N}), so that φΨCc(2×+×𝒰)\varphi\circ\Psi\in C^{\infty}_{c}({\mathbb{R}}^{2}\times{\mathbb{R}}_{+}^{*}\times{\mathcal{U}}). Thanks to (6) and (8), we have

(9) Y(φ,φ)=122×+×𝕊I(z,r,u)dzrνdrβ(du),\displaystyle{\mathcal{E}}^{Y}(\varphi,\varphi)=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}I(z,r,u){\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u),

where

I(z,r,u)=12Nz(φΨ)(z,r,u)2+2r|r(φΨ)(z,r,u)|2+12r𝕊(φΨ)(z,r,u)2.\displaystyle I(z,r,u)=\frac{1}{2N}||\nabla_{z}(\varphi\circ\Psi)(z,r,u)||^{2}+2r|\partial_{r}(\varphi\circ\Psi)(z,r,u)|^{2}+\frac{1}{2r}||\nabla_{\mathbb{S}}(\varphi\circ\Psi)(z,r,u)||^{2}.

We recall that for φ:(2)N\varphi:({\mathbb{R}}^{2})^{N}\to{\mathbb{R}}, we call φ(x)=((φ(x))1,,(φ(x))N)(2)N\nabla\varphi(x)=((\nabla\varphi(x))^{1},\dots,(\nabla\varphi(x))^{N})\in({\mathbb{R}}^{2})^{N} the total gradient of φ\varphi at x(2)Nx\in({\mathbb{R}}^{2})^{N}, and we have (φ(x))i2(\nabla\varphi(x))^{i}\in{\mathbb{R}}^{2} for each i[[1,N]]i\in[\![1,N]\!]. And for ϕ:Op\phi:O\to{\mathbb{R}}^{p}, where OO is open in n{\mathbb{R}}^{n}, we denote by dzϕ{\rm d}_{z}\phi the differential of ϕ\phi at zOz\in O.

We start with the study of Ψ(z,r,u)=γ(z)+ru\Psi(z,r,u)=\gamma(z)+\sqrt{r}u, where we recall that γ\gamma was introduced in Section 2 and that Φ𝕊(x)=πHx/πHx\Phi_{\mathbb{S}}(x)=\pi_{H}x/||\pi_{H}x|| is defined on a neighborhood of 𝕊{\mathbb{S}} in (2)N({\mathbb{R}}^{2})^{N}, see (3). It holds that for all (z,r,u)2×+×𝕊(z,r,u)\in{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}^{*}\times{\mathbb{S}} and all h2h\in{\mathbb{R}}^{2}, kk\in{\mathbb{R}} and (2)N\ell\in({\mathbb{R}}^{2})^{N},

dzΨ(,r,u)(h)=γ(h),drΨ(z,,u)(k)=k2ru,du[Ψ(z,r,Φ𝕊())]()=rπu(πH()),\displaystyle{\rm d}_{z}\Psi(\cdot,r,u)(h)=\gamma(h),\qquad{\rm d}_{r}\Psi(z,\cdot,u)(k)=\frac{k}{2\sqrt{r}}u,\qquad{\rm d}_{u}[\Psi(z,r,\Phi_{\mathbb{S}}(\cdot))](\ell)=\sqrt{r}\pi_{u^{\perp}}(\pi_{H}(\ell)),

For the first equality, it suffices to use that γ\gamma is linear, so that dzΨ(,r,u)(h)=dzγ(h)=γ(h){\rm d}_{z}\Psi(\cdot,r,u)(h)={\rm d}_{z}\gamma(h)=\gamma(h). The second equality is obvious. For the third equality, which is the differential at u𝕊u\in{\mathbb{S}} of the function F(x)=γ(z)+rΦ𝕊(x)F(x)=\gamma(z)+\sqrt{r}\Phi_{\mathbb{S}}(x) defined for xENx\in E_{N} (which is open in (2)N({\mathbb{R}}^{2})^{N} and contains 𝕊{\mathbb{S}}), we write duF=rduΦ𝕊{\rm d}_{u}F=\sqrt{r}{\rm d}_{u}\Phi_{\mathbb{S}}. But ΦS=GπH\Phi_{S}=G\circ\pi_{H}, where G(x)=x/xG(x)=x/||x||, and we have duπH=πH{\rm d}_{u}\pi_{H}=\pi_{H} and dπH(u)G=duG=πu{\rm d}_{\pi_{H}(u)}G={\rm d}_{u}G=\pi_{u^{\perp}} for u𝕊u\in{\mathbb{S}}. All in all, duF=rπuπH{\rm d}_{u}F=\sqrt{r}\pi_{u^{\perp}}\circ\pi_{H}.

First, we have z(φΨ)(z,r,u)=i=1N[φ(Ψ(z,r,u))]i\nabla_{z}(\varphi\circ\Psi)(z,r,u)=\sum_{i=1}^{N}[\nabla\varphi(\Psi(z,r,u))]^{i}. Indeed, for all h2h\in{\mathbb{R}}^{2}, it holds that

dz(φΨ(,r,u))(h)=(dΨ(z,r,u)φ)[(dzΨ(,r,u))(h)]=(dΨ(z,r,u)φ)(γ(h))=φ(Ψ(z,r,u))γ(h),{\rm d}_{z}(\varphi\circ\Psi(\cdot,r,u))(h)=({\rm d}_{\Psi(z,r,u)}\varphi)[({\rm d}_{z}\Psi(\cdot,r,u))(h)]=({\rm d}_{\Psi(z,r,u)}\varphi)(\gamma(h))=\nabla\varphi(\Psi(z,r,u))\cdot\gamma(h),

which, by definition of γ\gamma, equals hi=1N[φ(Ψ(z,r,u))]ih\cdot\sum_{i=1}^{N}[\nabla\varphi(\Psi(z,r,u))]^{i}.

This implies that

(10) 12Nz(φΨ(z,r,u))2=12Ni=1N[φ(Ψ(z,r,u))]i2=12πH(φ(Ψ(z,r,u)))2.\frac{1}{2N}\|\nabla_{z}(\varphi\circ\Psi(z,r,u))\|^{2}=\frac{1}{2N}\Big{\|}\sum_{i=1}^{N}[\nabla\varphi(\Psi(z,r,u))]^{i}\Big{\|}^{2}=\frac{1}{2}\|\pi_{H^{\perp}}(\nabla\varphi(\Psi(z,r,u)))\|^{2}.

Indeed, recalling the expression of πH\pi_{H}, see Section 2, it suffices to note that for all x(2)Nx\in({\mathbb{R}}^{2})^{N}, πH(x)2=γ(S[[1,N]](x))2=NS[[1,N]](x)2=N1i=1Nxi2\|\pi_{H^{\perp}}(x)\|^{2}=\|\gamma(S_{[\![1,N]\!]}(x))\|^{2}=N\|S_{[\![1,N]\!]}(x)\|^{2}=N^{-1}\|\sum_{i=1}^{N}x^{i}\|^{2}.

Next, r(φΨ)(z,r,u)=(φ)(Ψ(z,r,u))u/(2r)\partial_{r}(\varphi\circ\Psi)(z,r,u)=(\nabla\varphi)(\Psi(z,r,u))\cdot u/(2\sqrt{r}). Indeed, for kk\in{\mathbb{R}},

dr(φΨ(z,,u))(k)=(dΨ(z,r,u)φ)[(drΨ(z,,u))(k)]=(dΨ(z,r,u)φ)(u)×k2r,{\rm d}_{r}(\varphi\circ\Psi(z,\cdot,u))(k)=({\rm d}_{\Psi(z,r,u)}\varphi)[({\rm d}_{r}\Psi(z,\cdot,u))(k)]=({\rm d}_{\Psi(z,r,u)}\varphi)(u)\times\frac{k}{2\sqrt{r}},

which is nothing but (φ)(Ψ(z,r,u))u×k/(2r)(\nabla\varphi)(\Psi(z,r,u))\cdot u\times k/(2\sqrt{r}).

This implies, recalling that πu\pi_{u} is the orthogonal projection on Span(u)(2)N{\rm Span}(u)\subset({\mathbb{R}}^{2})^{N}, that

(11) 2r|r(φΨ)(z,r,u)|2=12πu((φ)(Ψ(z,r,u)))2=12πH(πu((φ)(Ψ(z,r,u))))22r|\partial_{r}(\varphi\circ\Psi)(z,r,u)|^{2}=\frac{1}{2}\|\pi_{u}((\nabla\varphi)(\Psi(z,r,u)))\|^{2}=\frac{1}{2}\|\pi_{H}(\pi_{u}((\nabla\varphi)(\Psi(z,r,u))))\|^{2}

since u𝕊u\in{\mathbb{S}}, so that u=1||u||=1 and uHu\in H.

Finally, 𝕊(φΨ)(z,r,u)=rπH(πu(φ(Ψ(z,r,u))))\nabla_{\mathbb{S}}(\varphi\circ\Psi)(z,r,u)=\sqrt{r}\pi_{H}(\pi_{u^{\perp}}(\nabla\varphi(\Psi(z,r,u)))). Indeed, for all (2)N\ell\in({\mathbb{R}}^{2})^{N},

du((φΨ)(z,r,Φ𝕊()))()=\displaystyle d_{u}((\varphi\circ\Psi)(z,r,\Phi_{\mathbb{S}}(\cdot)))(\ell)= (dΨ(z,r,u)φ)(du[Ψ(z,r,Φ𝕊())]())\displaystyle(d_{\Psi(z,r,u)}\varphi)(d_{u}[\Psi(z,r,\Phi_{\mathbb{S}}(\cdot))](\ell))
=\displaystyle= r(dΨ(z,r,u)φ)(πu(πH()))\displaystyle\sqrt{r}(d_{\Psi(z,r,u)}\varphi)(\pi_{u^{\perp}}(\pi_{H}(\ell)))
=\displaystyle= rφ(Ψ(z,r,u))πu(πH())\displaystyle\sqrt{r}\nabla\varphi(\Psi(z,r,u))\cdot\pi_{u^{\perp}}(\pi_{H}(\ell))
=\displaystyle= rπH(πu(φ(Ψ(z,r,u)))),\displaystyle\sqrt{r}\pi_{H}(\pi_{u^{\perp}}(\nabla\varphi(\Psi(z,r,u))))\cdot\ell,

and we conclude since 𝕊(φΨ)(z,r,u)=x((φΨ)(z,r,Φ𝕊()))(u)\nabla_{\mathbb{S}}(\varphi\circ\Psi)(z,r,u)=\nabla_{x}((\varphi\circ\Psi)(z,r,\Phi_{\mathbb{S}}(\cdot)))(u) by definition of 𝕊\nabla_{\mathbb{S}}, see (5).

This implies that

(12) 12r𝕊(φΨ)(z,r,u)2=12πH(πu(φ(Ψ(z,r,u))))2.\frac{1}{2r}||\nabla_{\mathbb{S}}(\varphi\circ\Psi)(z,r,u)||^{2}=\frac{1}{2}\|\pi_{H}(\pi_{u^{\perp}}(\nabla\varphi(\Psi(z,r,u))))\|^{2}.

Gathering (10), (11) and (12), we see that I(z,r,u)=12φ(Ψ(z,r,u))2I(z,r,u)=\frac{1}{2}\|\nabla\varphi(\Psi(z,r,u))\|^{2}, since for x(2)Nx\in({\mathbb{R}}^{2})^{N},

πH(x)2+πH(πu(x))2+πH(πu(x))2=x2\|\pi_{H^{\perp}}(x)\|^{2}+\|\pi_{H}(\pi_{u}(x))\|^{2}+\|\pi_{H}(\pi_{u^{\perp}}(x))\|^{2}=\|x\|^{2}

because u𝕊Hu\in{\mathbb{S}}\subset H.

Injecting the value of II in (9) and using Lemma A.2, we obtain

Y(φ,φ)=142×+×𝕊φ(Ψ(z,r,u))2dzrνdrβ(du)=12(2)Nφ2dμ.\displaystyle{\mathcal{E}}^{Y}(\varphi,\varphi)=\frac{1}{4}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}^{*}\times{\mathbb{S}}}\|\nabla\varphi(\Psi(z,r,u))\|^{2}{\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\|\nabla\varphi\|^{2}{\rm d}\mu.

Step 5. As a last technical step, we verify that (Y,Y)({\mathcal{E}}^{Y},{\mathcal{F}}^{Y}) is a regular Dirichlet space on L2((2)N,μ)L^{2}(({\mathbb{R}}^{2})^{N},\mu) with core Cc(𝒳EN)C_{c}^{\infty}({\mathcal{X}}\cap E_{N}), i.e. that for all φY\varphi\in{\mathcal{F}}^{Y}, there is φnCc(𝒳EN)\varphi_{n}\in C_{c}^{\infty}({\mathcal{X}}\cap E_{N}) such that limnφnφL2((2)N,μ)+Y(φnφ,φnφ)=0\lim_{n}||\varphi_{n}-\varphi||_{L^{2}(({\mathbb{R}}^{2})^{N},\mu)}+{\mathcal{E}}^{Y}(\varphi_{n}-\varphi,\varphi_{n}-\varphi)=0.

Recalling (7) and using that ((M,D,UA),(M,D,UA))({\mathcal{E}}^{(M,D,U_{A})},{\mathcal{F}}^{(M,D,U_{A})}) on L2(2×+×𝕊,dzrνdrβ(du))L^{2}({\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}},{\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u)) is regular with core Cc(2×+×𝒰)C_{c}^{\infty}({\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathcal{U}}), there is gnCc(2×+×𝒰)g_{n}\in C_{c}^{\infty}({\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathcal{U}}) such that

gnφΨL2(2×+×𝕊,dzrνdrβ(du))+(M,D,UA)(gnφΨ,gnφΨ)0.||g_{n}-\varphi\circ\Psi||_{L^{2}({\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}},{\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u))}+{\mathcal{E}}^{(M,D,U_{A})}(g_{n}-\varphi\circ\Psi,g_{n}-\varphi\circ\Psi)\to 0.

Setting φn=gnΨ1\varphi_{n}=g_{n}\circ\Psi^{-1}, it holds that φnCc(𝒳EN)\varphi_{n}\in C_{c}^{\infty}({\mathcal{X}}\cap E_{N}) and we have, by (8),

Y(φnφ,φnφ)=12(M,D,UA)(gnφΨ,gnφΨ)0,{\mathcal{E}}^{Y}(\varphi_{n}-\varphi,\varphi_{n}-\varphi)=\frac{1}{2}{\mathcal{E}}^{(M,D,U_{A})}(g_{n}-\varphi\circ\Psi,g_{n}-\varphi\circ\Psi)\to 0,

as well as, by Lemma A.2,

φnφL2((2)N,μ)=12gnφΨL2(2×+×𝕊,dzrνdrβ(du))0.||\varphi_{n}-\varphi||_{L^{2}(({\mathbb{R}}^{2})^{N},\mu)}=\frac{1}{2}||g_{n}-\varphi\circ\Psi||_{L^{2}({\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}},{\rm d}zr^{\nu}{\rm d}r\beta({\rm d}u))}\to 0.

Step 6. By Steps 3, 4 and 5, we know that 𝕐{\mathbb{Y}} is a μ\mu-symmetric (𝒳EN){}({\mathcal{X}}\cap E_{N})\cup\{\triangle\}-valued diffusion with regular Dirichlet space (Y,Y)({\mathcal{E}}^{Y},{\mathcal{F}}^{Y}) with core Cc(𝒳EN)C_{c}^{\infty}({\mathcal{X}}\cap E_{N}) and with Y(φ,φ)=12(2)Nφ2dμ{\mathcal{E}}^{Y}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}||\nabla\varphi||^{2}{\rm d}\mu for φCc(𝒳EN)\varphi\in C_{c}^{\infty}({\mathcal{X}}\cap E_{N}).

Now, applying Lemma B.6 to 𝕏\mathbb{X} defined in Proposition 4.1-(i) with the open set 𝒳EN{\mathcal{X}}\cap E_{N}, we see that 𝕏{\mathbb{X}^{*}}, i.e. 𝕏{\mathbb{X}} killed when getting outside 𝒳EN{\mathcal{X}}\cap E_{N}, is a μ\mu-symmetric (𝒳EN){}({\mathcal{X}}\cap E_{N})\cup\{\triangle\}-valued diffusion process with regular Dirichlet space (X,X)({\mathcal{E}}^{X^{*}},{\mathcal{F}}^{X^{*}}) with core Cc(𝒳EN)C_{c}^{\infty}({\mathcal{X}}\cap E_{N}) and with X(φ,φ)=12(2)Nφ2dμ{\mathcal{E}}^{X^{*}}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}||\nabla\varphi||^{2}{\rm d}\mu for φCc(𝒳EN)\varphi\in C_{c}^{\infty}({\mathcal{X}}\cap E_{N}).

This implies, as recalled in Subsection B.1, that (X,X)=(Y,Y)({\mathcal{E}}^{X^{*}},{\mathcal{F}}^{X^{*}})=({\mathcal{E}}^{Y},{\mathcal{F}}^{Y}). The conclusion follows by uniqueness, see [11, Theorem 4.2.8 p 167]. ∎

Actually, (R[[1,N]](Xt))t0(R_{[\![1,N]\!]}(X_{t}))_{t\geq 0} and (S[[1,N]](Xt))t0(S_{[\![1,N]\!]}(X_{t}))_{t\geq 0} are some independent squared Bessel process and Brownian motion until explosion (and not only until the first time where R[[1,N]](Xt)=0R_{[\![1,N]\!]}(X_{t})=0, as shown in Proposition 5.1), a fact that we shall often use.

Lemma 5.2.

We fix N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta and we consider a QKS(θ,N)QKS(\theta,N)-process 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳){\mathbb{X}}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}_{x}^{X})_{x\in{\mathcal{X}_{\triangle}}}). Quasi-everywhere, there are a 2D2D-Brownian motion (Mt)t0(M_{t})_{t\geq 0} with diffusion constant N1/2N^{-1/2} issued from S[[1,N]](x)S_{[\![1,N]\!]}(x) and a squared Bessel process (Dt)t0(D_{t})_{t\geq 0} with dimension dθ,N(N)d_{\theta,N}(N) issued from R[[1,N]](x)R_{[\![1,N]\!]}(x) (killed when it gets out of (0,)(0,\infty) if dθ,N(N)0d_{\theta,N}(N)\leq 0) independent of (Mt)t0(M_{t})_{t\geq 0} such that xX{\mathbb{P}}_{x}^{X}-a.s., S[[1,N]](Xt)=MtS_{[\![1,N]\!]}(X_{t})=M_{t} and R[[1,N]](Xt)=DtR_{[\![1,N]\!]}(X_{t})=D_{t} for all t[0,ζ)t\in[0,\zeta).

Proof.

If θ2\theta\geq 2, this follows from Proposition 5.1: setting τ=inf{t>0:R[[1,N]](Xt)(0,)}\tau=\inf\{t>0:R_{[\![1,N]\!]}(X_{t})\notin(0,\infty)\}, we have τ=ζ\tau=\zeta. Indeed, on {τ<ζ}\{\tau<\zeta\}, we have XτENX_{\tau}\notin E_{N}, whence Xτ𝒳X_{\tau}\notin{\mathcal{X}} since 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} with k0Nk_{0}\leq N (because θ2\theta\geq 2), which contradicts the fact that τ<ζ\tau<\zeta.

We now suppose that θ<2\theta<2, so that k0>Nk_{0}>N and thus 𝒳=(2)N{\mathcal{X}}=({\mathbb{R}}^{2})^{N}. We introduce the shortened notation R(x)=R[[1,N]](x)R(x)=R_{[\![1,N]\!]}(x), S(x)=S[[1,N]](x)S(x)=S_{[\![1,N]\!]}(x) and split the proof in three parts.

Step 1. First, one can show similarly (but much more easily) as in the proof of Proposition 5.1 that there exists a 2D2D-Brownian motion (Mt)t0(M_{t})_{t\geq 0} independent of (Xtγ(S(Xt)))t0(X_{t}-\gamma(S(X_{t})))_{t\geq 0}, such that S(Xt)=MtS(X_{t})=M_{t} for all t[0,ζ)t\in[0,\zeta). This moreover shows that (Mt)t0(M_{t})_{t\geq 0} is independent of (R(Xt))t0(R(X_{t}))_{t\geq 0}, because R(Xt)=Xtγ(S(Xt))2R(X_{t})=\|X_{t}-\gamma(S(X_{t}))\|^{2}.

Step 2. We consider some function gmCc((2)N)g_{m}\in C^{\infty}_{c}(({\mathbb{R}}^{2})^{N}) such that gm=1g_{m}=1 on B(0,m)B(0,m) and supα(0,1]supx(2)N|αXgm(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}g_{m}(x)|<\infty. Such a function exists by Remark 6.3. For φCc(+)\varphi\in C^{\infty}_{c}({\mathbb{R}}_{+}), we set ψ(x)=φ(R(x))\psi(x)=\varphi(R(x)) and show that ψgm𝒟𝒜X\psi g_{m}\in{\mathcal{D}}_{{\mathcal{A}}^{X}} and that for all xB(0,m)x\in B(0,m),

(13) 𝒜X(ψgm)(x)=\displaystyle{\mathcal{A}}^{X}(\psi g_{m})(x)= 2R(x)φ′′(R(x))+dθ,N(N)φ(R(x)).\displaystyle 2R(x)\varphi^{\prime\prime}(R(x))+d_{\theta,N}(N)\varphi^{\prime}(R(x)).

To this end, we apply Remark 4.3. Since ψgmCc((2)N)\psi g_{m}\in C^{\infty}_{c}(({\mathbb{R}}^{2})^{N}) and since 𝒳=(2)N{\mathcal{X}}=({\mathbb{R}}^{2})^{N}, we have to show that supα(0,1]supx(2)N|αX(ψgm)(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}(\psi g_{m})(x)|<\infty, and we will deduce that 𝒜X(ψgm)=X(ψgm){\mathcal{A}}^{X}(\psi g_{m})={\mathcal{L}}^{X}(\psi g_{m}). By (3), we have αX(ψgm)=gmαXψ+ψαXgm+ψgm{\mathcal{L}}^{X}_{\alpha}(\psi g_{m})=g_{m}{\mathcal{L}}^{X}_{\alpha}\psi+\psi{\mathcal{L}}^{X}_{\alpha}g_{m}+\nabla\psi\cdot\nabla g_{m}. The only difficulty consists in showing that supα(0,1]supx(2)N|αXψ(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}\psi(x)|<\infty. Using that xiR(x)=2(xiS(x))\nabla_{x^{i}}R(x)=2(x^{i}-S(x)), we find xiψ(x)=2(xiS(x))φ(R(x)).\nabla_{x^{i}}\psi(x)=2(x^{i}-S(x))\varphi^{\prime}(R(x)). Hence by symmetry,

θN1ijNxixjxixj2+αxiψ(x)=\displaystyle\frac{\theta}{N}\sum_{1\leq i\neq j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot\nabla_{x^{i}}\psi(x)= 2θNφ(R(x))1ijNxixjxixj2+αxi\displaystyle\frac{2\theta}{N}\varphi^{\prime}(R(x))\sum_{1\leq i\neq j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot x^{i}
(14) =\displaystyle= θNφ(R(x))1ijNxixj2xixj2+α.\displaystyle\frac{\theta}{N}\varphi^{\prime}(R(x))\sum_{1\leq i\neq j\leq N}\frac{\|x^{i}-x^{j}\|^{2}}{\|x^{i}-x^{j}\|^{2}+\alpha}.

Besides, Δxiψ(x)=4(11/N)φ(R(x))+4xiS(x)2φ′′(R(x)),\Delta_{x^{i}}\psi(x)=4(1-1/N)\varphi^{\prime}(R(x))+4\|x^{i}-S(x)\|^{2}\varphi^{\prime\prime}(R(x)), whence

(15) Δψ(x)=\displaystyle\Delta\psi(x)= 4(N1)φ(R(x))+4R(x)φ′′(R(x)).\displaystyle 4(N-1)\varphi^{\prime}(R(x))+4R(x)\varphi^{\prime\prime}(R(x)).

We conclude by combining (5) and (15) that

αXψ(x)=\displaystyle{\mathcal{L}}^{X}_{\alpha}\psi(x)= 2R(x)φ′′(R(x))+(2(N1)θN1ijNxixj2xixj2+α)φ(R(x)).\displaystyle 2R(x)\varphi^{\prime\prime}(R(x))+\Big{(}2(N-1)-\frac{\theta}{N}\sum_{1\leq i\neq j\leq N}\frac{\|x^{i}-x^{j}\|^{2}}{\|x^{i}-x^{j}\|^{2}+\alpha}\Big{)}\varphi^{\prime}(R(x)).

We immediately deduce, since φ\varphi is compactly supported, that supα(0,1]supx(2)N|αXψ(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}\psi(x)|<\infty, whence supα(0,1]supx(2)N|αX(ψgm)(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}(\psi g_{m})(x)|<\infty. Hence ψgm𝒟𝒜X\psi g_{m}\in{\mathcal{D}}_{{\mathcal{A}}^{X}} and 𝒜X(ψgm)=X(ψgm){\mathcal{A}}^{X}(\psi g_{m})={\mathcal{L}}^{X}(\psi g_{m}). Moreover, recalling that Xψ=αXψ{\mathcal{L}}^{X}\psi={\mathcal{L}}^{X}_{\alpha}\psi with α=0\alpha=0 and that gm=1g_{m}=1 on B(0,m)B(0,m), we conclude that 𝒜X(ψgm)(x)=0Xψ(x){\mathcal{A}}^{X}(\psi g_{m})(x)={\mathcal{L}}^{X}_{0}\psi(x) for xB(0,m)x\in B(0,m), whence (13), because 2(N1)θ(N1)=dθ,N(N)2(N-1)-\theta(N-1)=d_{\theta,N}(N).

Step 3. We define ζm=inf{t>0:XtB(0,m)}\zeta_{m}=\inf\{t>0:X_{t}\notin B(0,m)\}. By Lemma B.2 and Step 1, for all φCc(+)\varphi\in C^{\infty}_{c}({\mathbb{R}}_{+}), quasi-everywhere in B(0,m)B(0,m), φ(R(Xtζm))φ(R(x))0tζmXφ(R(Xs))ds\varphi(R(X_{t\land\zeta_{m}}))-\varphi(R(x))-\int_{0}^{t\land\zeta_{m}}{\mathcal{L}}^{X}\varphi(R(X_{s})){\rm d}s is a xX{\mathbb{P}}_{x}^{X}-martingale. Recalling (13), we classically conclude that there is a Brownian motion WW such that R(Xt)=R(x)+20tR(Xs)dWs+dθ,N(N)tR(X_{t})=R(x)+2\int_{0}^{t}\sqrt{R(X_{s})}{\rm d}W_{s}+d_{\theta,N}(N)t during [0,ζn][0,\zeta_{n}]. We recognize the S.D.E. of a squared Bessel process with dimension dθ,N(N)d_{\theta,N}(N), see Revuz-Yor [21, Chapter XI]. Since we know from Remark 4.2 that ζ=limmζm\zeta=\lim_{m}\zeta_{m}, the proof is complete. ∎

6. Some cutoff functions

We will need several times to approximate some indicator functions by some smooth functions, on which the generator X{\mathcal{L}}^{X} (or U{\mathcal{L}}^{U}) is bounded. This does not seem obvious, due to the singularity of X{\mathcal{L}}^{X}. We recall that αX{\mathcal{L}}^{X}_{\alpha} and αU{\mathcal{L}}^{U}_{\alpha} were defined in (2) and (4).

Lemma 6.1.

Fix N2N\geq 2, θ>0\theta>0, recall that k0=2N/θk_{0}=\lceil 2N/\theta\rceil and that 𝒳=Ek0{\mathcal{X}}=E_{k_{0}}. Consider a partition 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} and define, for ε[0,1]\varepsilon\in[0,1], (with the convention that B(0,1/0)=(2)NB(0,1/0)=({\mathbb{R}}^{2})^{N}),

G𝐊,ε={x𝒳:min1pqminiKp,jKqxixj2>ε}B(0,1ε).G_{{\mathbf{K}},\varepsilon}=\Big{\{}x\in{\mathcal{X}}:\min_{1\leq p\neq q\leq\ell}\;\;\min_{i\in K_{p},j\in K_{q}}||x^{i}-x^{j}||^{2}>\varepsilon\Big{\}}\cap B\Big{(}0,\frac{1}{\varepsilon}\Big{)}.

(i) For all ε(0,1]\varepsilon\in(0,1], there is a family of open relatively compact subsets G𝐊,εnG^{n}_{{\mathbf{K}},\varepsilon} of G𝐊,0G_{{\mathbf{K}},0} such that

n1G𝐊,εnG𝐊,εand for each n1G𝐊,εnG𝐊,εn+1,\bigcup_{n\geq 1}G^{n}_{{\mathbf{K}},\varepsilon}\supset G_{{\mathbf{K}},\varepsilon}\quad\hbox{and for each $n\geq 1$, }G^{n}_{{\mathbf{K}},\varepsilon}\subset G^{n+1}_{{\mathbf{K}},\varepsilon},

and some of [0,1][0,1]-valued functions Γ𝐊,εnCc(G𝐊,0)\Gamma_{{\mathbf{K}},\varepsilon}^{n}\in C^{\infty}_{c}(G_{{\mathbf{K}},0}) such that for some η(0,1]\eta\in(0,1], for all n1n\geq 1,

SuppΓ𝐊,εnG𝐊,η,Γ𝐊,εn=1 on G𝐊,εnandsupα(0,1]supx(2)N|αXΓ𝐊,εn(x)|<.{\rm Supp}\;\Gamma_{{\mathbf{K}},\varepsilon}^{n}\subset G_{{\mathbf{K}},\eta},\quad\Gamma_{{\mathbf{K}},\varepsilon}^{n}=1\;\hbox{ on }\;G_{{\mathbf{K}},\varepsilon}^{n}\quad\hbox{and}\quad\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}\Big{|}{\mathcal{L}}^{X}_{\alpha}\Gamma_{{\mathbf{K}},\varepsilon}^{n}(x)\Big{|}<\infty.

(ii) With the same sets G𝐊,εnG^{n}_{{\mathbf{K}},\varepsilon} as in (i), there is a family of functions Γ𝐊,ε𝕊,nCc(𝕊G𝐊,0)\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}\in C^{\infty}_{c}({\mathbb{S}}\cap G_{{\mathbf{K}},0}) with values in [0,1][0,1] such that for all n1n\geq 1,

Γ𝐊,ε𝕊,n=1 on 𝕊G𝐊,εnandsupα(0,1]supu𝕊|αUΓ𝐊,ε𝕊,n(u)|<.\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}=1\;\hbox{ on }\;{\mathbb{S}}\cap G^{n}_{{\mathbf{K}},\varepsilon}\quad\hbox{and}\quad\sup_{\alpha\in(0,1]}\sup_{u\in{\mathbb{S}}}\Big{|}{\mathcal{L}}^{U}_{\alpha}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u)\Big{|}<\infty.

The section is devoted to the proof of this lemma. We start with the following technical result.

Lemma 6.2.

We define the family (c)[[1,N]](c_{\ell})_{\ell\in[\![1,N]\!]} by c0=1c_{0}=1 and for all [[1,N1]]\ell\in[\![1,N-1]\!], c+1=(2+4)cc_{\ell+1}=\color[rgb]{0,0,0}(2+4\ell)c_{\ell}. For all K[[1,N]]K\subsetneq[\![1,N]\!], all ε(0,1]\varepsilon\in(0,1], all x(2)Nx\in({\mathbb{R}}^{2})^{N} such that

RK(x)2c|K|εandminjKRK{j}(x)c|K|+1ε,R_{K}(x)\leq 2c_{|K|}\varepsilon\quad\hbox{and}\quad\min_{j\notin K}R_{K\cup\{j\}}(x)\geq c_{|K|+1}\varepsilon,

it holds that xixj2c|K|ε\|x^{i}-x^{j}\|^{2}\geq c_{|K|}\varepsilon for all iKi\in K, all jKj\notin K.

Proof.

We fix K[[1,N]]K\subsetneq[\![1,N]\!], ε(0,1]\varepsilon\in(0,1] and x(2)Nx\in({\mathbb{R}}^{2})^{N} as in the statement and assume by contradiction that there are i0Ki_{0}\in K, j0Kj_{0}\notin K such that xi0xj02<c|K|ε\|x^{i_{0}}-x^{j_{0}}\|^{2}<c_{|K|}\varepsilon. Then for all iKi\in K,

xj0xi22xi0xj02+2xi0xi22xi0xj02+2|K|RK(x)<(2+4|K|)c|K|ε.\|x^{j_{0}}-x^{i}\|^{2}\leq 2\|x^{i_{0}}-x^{j_{0}}\|^{2}+2\|x^{i_{0}}-x^{i}\|^{2}\leq 2\|x^{i_{0}}-x^{j_{0}}\|^{2}+\color[rgb]{0,0,0}2|K|\color[rgb]{0,0,0}R_{K}(x)<\color[rgb]{0,0,0}(2+4|K|)\color[rgb]{0,0,0}c_{|K|}\varepsilon.

This implies that

RK{j0}(x)=12(|K|+1)(2|K|RK(x)+2iKxj0xi2)RK(x)+1|K|+1iKxj0xi2,R_{K\cup\{j_{0}\}}(x)=\frac{1}{2(|K|+1)}\Big{(}2|K|R_{K}(x)+2\sum_{i\in K}\|x^{j_{0}}-x^{i}\|^{2}\Big{)}\leq R_{K}(x)+\frac{1}{|K|+1}\sum_{i\in K}\|x^{j_{0}}-x^{i}\|^{2},

whence

RK{j0}(x)<2c|K|ε+2+4|K||K|+1|K|c|K|ε<(2+4|K|)c|K|ε=c|K|+1ε,R_{K\cup\{j_{0}\}}(x)<2c_{|K|}\varepsilon+\frac{2+4|K|}{|K|+1}|K|c_{|K|}\varepsilon<(2+4|K|)c_{|K|}\varepsilon=c_{|K|+1}\varepsilon,

which is a contradiction. ∎

We are now ready to give the

Proof of Lemma 6.1.

We introduce some nondecreasing CC^{\infty} function ϱ:+[0,1]\varrho:{\mathbb{R}}_{+}\to[0,1] such that ϱ=0\varrho=0 on [0,1/2][0,1/2] and ϱ=1\varrho=1 on [1,)[1,\infty). We divide the proof in three steps.

Step 1. We fix n1n\geq 1 and define, for K[[1,N]]K\subset[\![1,N]\!], using the family (c)[[1,N]](c_{\ell})_{\ell\in[\![1,N]\!]} of Lemma 6.2,

E~K,n={x(2)N:LK,RL(x)>c|L|n}andΓ~K,n(x)=LKϱ(nRL(x)c|L|),\tilde{E}_{K,n}=\Big{\{}x\in({\mathbb{R}}^{2})^{N}:\forall\;L\supset K,\;R_{L}(x)>\frac{c_{|L|}}{n}\Big{\}}\qquad\hbox{and}\qquad\tilde{\Gamma}_{K,n}(x)=\prod_{L\supset K}\varrho\Big{(}\frac{nR_{L}(x)}{c_{|L|}}\Big{)},

where {LK}={L[[1,N]]:KL}\{L\supset K\}=\{L\subset[\![1,N]\!]:K\subset L\}. We have

(1) Γ~K,nC((2)N),Supp Γ~K,nE~K,2nandΓ~K,n=1onE~K,n.\tilde{\Gamma}_{K,n}\in C^{\infty}(({\mathbb{R}}^{2})^{N}),\quad\mbox{Supp }\tilde{\Gamma}_{K,n}\subset\tilde{E}_{K,2n}\quad\mbox{and}\quad\tilde{\Gamma}_{K,n}=1\quad\mbox{on}\quad\tilde{E}_{K,n}.

Since RK(x)>0R_{K}(x)>0 implies that RL(x)>0R_{L}(x)>0 for all LKL\supset K, we also have

(2) n1E~K,n=E~K,whereE~K={x(2)N:RK(x)>0}.\cup_{n\geq 1}\tilde{E}_{K,n}=\tilde{E}_{K},\quad\hbox{where}\quad\tilde{E}_{K}=\{x\in({\mathbb{R}}^{2})^{N}:R_{K}(x)>0\}.

We now show, and this is the main difficulty of the step, that for all A>0A>0, all K[[1,N]]K\subset[\![1,N]\!] with |K|2|K|\geq 2, we have supα(0,1]supxB(0,A)|αXΓ~K,n(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in B(0,A)}|{\mathcal{L}}^{X}_{\alpha}\tilde{\Gamma}_{K,n}(x)|<\infty. Since supxB(0,A)|ΔΓ~K,n(x)|<\sup_{x\in B(0,A)}|\Delta\tilde{\Gamma}_{K,n}(x)|<\infty, we only have to verify that supα(0,1]supxB(0,A)|IK,n,α(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in B(0,A)}|I_{K,n,\alpha}(x)|<\infty, where

IK,n,α(x)=1ijNxixjxixj2xiΓ~K,n(x)=LKfK,L,n(x)1ijNxixjxixj2xiRL(x),I_{K,n,\alpha}(x)=\sum_{1\leq i\neq j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}}\cdot\nabla_{x^{i}}\tilde{\Gamma}_{K,n}(x)=\sum_{L\supset K}f_{K,L,n}(x)\sum_{1\leq i\neq j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}}\cdot\nabla_{x^{i}}R_{L}(x),

with

fK,L,n(x)=nc|L|ϱ(nRL(x)c|L|)MK,MLϱ(nRM(x)c|M|).f_{K,L,n}(x)=\frac{n}{c_{|L|}}\varrho^{\prime}\Big{(}\frac{nR_{L}(x)}{c_{|L|}}\Big{)}\prod_{M\supset K,M\neq L}\varrho\Big{(}\frac{nR_{M}(x)}{c_{|M|}}\Big{)}.

Using that xiRL(x)=2(xiSL(x))1I{iL}\nabla_{x^{i}}R_{L}(x)=2(x^{i}-S_{L}(x))\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{i\in L\}}, we now write

IK,n,α(x)=2LKfK,L,n(x)(AL,α(x)+BL,α(x)),I_{K,n,\alpha}(x)=2\sum_{L\supset K}f_{K,L,n}(x)(A_{L,\alpha}(x)+B_{L,\alpha}(x)),

where,

AL,α(x)=i,jL,ij(xixj)(xiSL(x))xixj2+αandBL,α(x)=iL,jLc(xixj)(xiSL(x))xixj2+α.A_{L,\alpha}(x)=\sum_{i,j\in L,i\neq j}\frac{(x^{i}-x^{j})\cdot(x^{i}-S_{L}(x))}{\|x^{i}-x^{j}\|^{2}+\alpha}\quad\hbox{and}\quad B_{L,\alpha}(x)=\sum_{i\in L,j\in L^{c}}\frac{(x^{i}-x^{j})\cdot(x^{i}-S_{L}(x))}{\|x^{i}-x^{j}\|^{2}+\alpha}.

We have supα(0,1]supxB(0,A)|fK,L,n(x)AL,α(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in B(0,A)}|f_{K,L,n}(x)A_{L,\alpha}(x)|<\infty because fK,L,nf_{K,L,n} is bounded and because

AL,α(x)=i,jL,ij(xixj)xixixj2+α=\displaystyle A_{L,\alpha}(x)=\sum_{i,j\in L,i\neq j}\frac{(x^{i}-x^{j})\cdot x^{i}}{\|x^{i}-x^{j}\|^{2}+\alpha}= 12i,jL,ijxixj2xixj2+α[0,|L|(|L|1)2].\displaystyle\frac{1}{2}\sum_{i,j\in L,i\neq j}\frac{\|x^{i}-x^{j}\|^{2}}{\|x^{i}-x^{j}\|^{2}+\alpha}\in\Big{[}0,\frac{|L|(|L|-1)}{2}\Big{]}.

Next, we assume that L[[1,N]]L\subsetneq[\![1,N]\!] (else BL,α(x)=0B_{L,\alpha}(x)=0) and observe that fK,L,n(x)0f_{K,L,n}(x)\neq 0 implies that RL(x)<c|L|/nR_{L}(x)<c_{|L|}/n (because ϱ=0\varrho^{\prime}=0 on [1,)[1,\infty)) and that miniLRL{i}(x)>c|L|+1/(2n)\min_{i\notin L}R_{L\cup\{i\}}(x)>c_{|L|+1}/(2n) (because ϱ=0\varrho=0 on [0,1/2][0,1/2]). By Lemma 6.2, this implies that miniL,jLcxixj2c|L|/(2n)\min_{i\in L,j\in L^{c}}||x^{i}-x^{j}||^{2}\geq c_{|L|}/(2n). We immediately conclude that supα(0,1]supxB(0,A)|fK,L,n(x)BL,α(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in B(0,A)}|f_{K,L,n}(x)B_{L,\alpha}(x)|<\infty.

Step 2. We can now prove (i). We fix ε(0,1]\varepsilon\in(0,1] and a partition 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} of [[1,N]][\![1,N]\!]. For some m1m\geq 1 to be chosen later (as a function of ε\varepsilon), for each n1n\geq 1, we set

G𝐊,εn=B(0,m)(K[[1,N]]:|K|=k0E~K,n)(1pqiKp,jKqE~{i,j},m),\displaystyle G_{{\mathbf{K}},\varepsilon}^{n}=B(0,m)\cap\Big{(}\bigcap_{K\subset[\![1,N]\!]:|K|=k_{0}}\tilde{E}_{K,n}\Big{)}\cap\Big{(}\bigcap_{1\leq p\neq q\leq\ell}\;\;\bigcap_{i\in K_{p},j\in K_{q}}\tilde{E}_{\{i,j\},m}\Big{)},
Γ𝐊,εn(x)=gm(x)(K[[1,N]]:|K|=k0Γ~K,n(x))(1pqiKp,jKqΓ~{i,j},m(x)),\displaystyle\Gamma_{{\mathbf{K}},\varepsilon}^{n}(x)=g_{m}(x)\Big{(}\prod_{K\subset[\![1,N]\!]:|K|=k_{0}}\tilde{\Gamma}_{K,n}(x)\Big{)}\Big{(}\prod_{1\leq p\neq q\leq\ell}\;\;\prod_{i\in K_{p},j\in K_{q}}\tilde{\Gamma}_{\{i,j\},m}(x)\Big{)},

where gm(x)=ϱ(m/x)g_{m}(x)=\varrho(m/\|x\|) with the extension gm(0)=1g_{m}(0)=1.

First, G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n} is clearly included in G𝐊,εn+1G_{{\mathbf{K}},\varepsilon}^{n+1} and relatively compact in G𝐊,0G_{{\mathbf{K}},0}. We deduce from (2) that, setting H𝐊,m=B(0,m)(1pqiKp,jKqE~{i,j},m)H_{{\mathbf{K}},m}=B(0,m)\cap(\cap_{1\leq p\neq q\leq\ell}\;\;\cap_{i\in K_{p},j\in K_{q}}\tilde{E}_{\{i,j\},m}),

n1G𝐊,εn=(K[[1,N]]:|K|=k0E~K)H𝐊,m=Ek0H𝐊,m=𝒳H𝐊,m.\bigcup_{n\geq 1}G_{{\mathbf{K}},\varepsilon}^{n}=\Big{(}\bigcap_{K\subset[\![1,N]\!]:|K|=k_{0}}\tilde{E}_{K}\Big{)}\cap H_{{\mathbf{K}},m}=E_{k_{0}}\cap H_{{\mathbf{K}},m}={\mathcal{X}}\cap H_{{\mathbf{K}},m}.

By (2) again, we can choose mm large enough so that H𝐊,mH_{{\mathbf{K}},m} contains G𝐊,εG_{{\mathbf{K}},\varepsilon}. Next, by (1), it holds that Γ𝐊,εnC((2)N)\Gamma_{{\mathbf{K}},\varepsilon}^{n}\in C^{\infty}(({\mathbb{R}}^{2})^{N}), that Γ𝐊,εn=1\Gamma_{{\mathbf{K}},\varepsilon}^{n}=1 on G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n} and that

SuppΓ𝐊,εnB(0,2m)(K[[1,N]]:|K|=k0E~K,2n)(1pqiKp,jKqE~{i,j},2m),{\rm Supp}\;\Gamma_{{\mathbf{K}},\varepsilon}^{n}\subset B(0,2m)\cap\Big{(}\bigcap_{K\subset[\![1,N]\!]:|K|=k_{0}}\tilde{E}_{K,2n}\Big{)}\cap\Big{(}\bigcap_{1\leq p\neq q\leq\ell}\;\;\bigcap_{i\in K_{p},j\in K_{q}}\tilde{E}_{\{i,j\},2m}\Big{)},

which is compact in G𝐊,0G_{{\mathbf{K}},0}. Moreover, SuppΓ𝐊,εnH𝐊,2m{\rm Supp}\;\Gamma_{{\mathbf{K}},\varepsilon}^{n}\subset H_{{\mathbf{K}},2m}. Since there exists η(0,1]\eta\in(0,1] such that H𝐊,2mG𝐊,ηH_{{\mathbf{K}},2m}\subset G_{{\mathbf{K}},\eta}, we conclude that Supp Γ𝐊,εnG𝐊,η\Gamma_{{\mathbf{K}},\varepsilon}^{n}\subset G_{{\mathbf{K}},\eta}.

It remains to show that supα(0,1]supx(2)N|αXΓ𝐊,εn(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}\Gamma^{n}_{{\mathbf{K}},\varepsilon}(x)|<\infty. Introducing

χ𝐊,εn(x)=(K[[1,N]]:|K|=k0Γ~K,n(x))(1pqiKp,jKqΓ~{i,j},m(x)),\chi_{{\mathbf{K}},\varepsilon}^{n}(x)=\Big{(}\prod_{K\subset[\![1,N]\!]:|K|=k_{0}}\tilde{\Gamma}_{K,n}(x)\Big{)}\Big{(}\prod_{1\leq p\neq q\leq\ell}\;\;\prod_{i\in K_{p},j\in K_{q}}\tilde{\Gamma}_{\{i,j\},m}(x)\Big{)},

which belongs to C((2)N)C^{\infty}(({\mathbb{R}}^{2})^{N}) by Step 1, we have Γ𝐊,εn=gmχ𝐊,εn(x)\Gamma^{n}_{{\mathbf{K}},\varepsilon}=g_{m}\chi_{{\mathbf{K}},\varepsilon}^{n}(x) (with the chosen value of mm) and thus by (3)

αXΓ𝐊,εn(x)=gm(x)αXχ𝐊,εn(x)+χ𝐊,εnαXgm(x)+gm(x)χ𝐊,εn(x).{\mathcal{L}}^{X}_{\alpha}\Gamma^{n}_{{\mathbf{K}},\varepsilon}(x)=g_{m}(x){\mathcal{L}}^{X}_{\alpha}\chi_{{\mathbf{K}},\varepsilon}^{n}(x)+\chi_{{\mathbf{K}},\varepsilon}^{n}{\mathcal{L}}^{X}_{\alpha}g_{m}(x)+\nabla g_{m}(x)\cdot\nabla\chi_{{\mathbf{K}},\varepsilon}^{n}(x).

The first term is uniformly bounded because gmg_{m} is bounded and supported in B(0,2m)B(0,2m) and because supα(0,1]supxB(0,2m)|Xχ𝐊,εn(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in B(0,2m)}|{\mathcal{L}}^{X}\chi_{{\mathbf{K}},\varepsilon}^{n}(x)|<\infty by Step 1 and (3). The third term is also uniformly bounded, since χ𝐊,εnC((2)N)\chi_{{\mathbf{K}},\varepsilon}^{n}\in C^{\infty}(({\mathbb{R}}^{2})^{N}) and since gm\nabla g_{m} is bounded and supported in B(0,2m)B(0,2m). Finally, the middle term is bounded because χ𝐊,εn\chi_{{\mathbf{K}},\varepsilon}^{n} is bounded by 11 and because αXgm{\mathcal{L}}^{X}_{\alpha}g_{m} is uniformly bounded, as we now show: Δgm\Delta g_{m} is obviously bounded since gmCc((2)N)g_{m}\in C^{\infty}_{c}(({\mathbb{R}}^{2})^{N}) and, since xigm(x)=mϱ(m/x)xi/x3\nabla_{x^{i}}g_{m}(x)=-m\varrho^{\prime}(m/||x||)x^{i}/\|x\|^{3},

1i,jNxixjxixj2+αxigm(x)=\displaystyle\sum_{1\leq i,j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot\nabla_{x^{i}}g_{m}(x)= mϱ(m/x)x31i,jNxixjxixj2+αxi\displaystyle-\frac{m\varrho^{\prime}(m/||x||)}{\|x\|^{3}}\sum_{1\leq i,j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot x^{i}
=\displaystyle= mϱ(m/x)2x31i,jNxixj2xixj2+α.\displaystyle-\frac{m\varrho^{\prime}(m/||x||)}{2\|x\|^{3}}\sum_{1\leq i,j\leq N}\frac{\|x^{i}-x^{j}\|^{2}}{\|x^{i}-x^{j}\|^{2}+\alpha}.

This last quantity is uniformly bounded, since ϱ\varrho^{\prime} is bounded and vanishes on [1,)[1,\infty).

Step 3. We now prove (ii), by showing that the restriction Γ𝐊,ε𝕊,n=Γ𝐊,εn|𝕊\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}=\Gamma_{{\mathbf{K}},\varepsilon}^{n}|_{\mathbb{S}} satisfies the required conditions. We obviously have Γ𝐊,ε𝕊,nCc(𝕊G𝐊,0)\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}\in C^{\infty}_{c}({\mathbb{S}}\cap G_{{\mathbf{K}},0}) and Γ𝐊,ε𝕊,n=1\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}=1 on 𝕊G𝐊,εn{\mathbb{S}}\cap G_{{\mathbf{K}},\varepsilon}^{n}. It remains to show that supα(0,1]supu𝕊|αUΓ𝐊,ε𝕊,n|<\sup_{\alpha\in(0,1]}\sup_{u\in{\mathbb{S}}}|{\mathcal{L}}^{U}_{\alpha}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}|<\infty, recall (4). Since Γ𝐊,ε𝕊,nC(𝕊)\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}\in C^{\infty}({\mathbb{S}}), Δ𝕊Γ𝐊,ε𝕊,n\Delta_{\mathbb{S}}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n} is bounded. We thus only have to verify that supα(0,1]supu𝕊|Tα(u)|<\sup_{\alpha\in(0,1]}\sup_{u\in{\mathbb{S}}}|T_{\alpha}(u)|<\infty, where

Tα(u)=θN1i,jNuiujuiuj2+α(𝕊Γ𝐊,ε𝕊,n(u))iT_{\alpha}(u)=-\frac{\theta}{N}\sum_{1\leq i,j\leq N}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(\nabla_{\mathbb{S}}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u))^{i}

Setting bαi(u)=θNj=1Nuiujuiuj2+αb^{i}_{\alpha}(u)=-\frac{\theta}{N}\sum_{j=1}^{N}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha} and using (7),

Tα(u)=bα(u)𝕊Γ𝐊,ε𝕊,n(u)=bα(u)πH(πu(Γ𝐊,ε𝕊,n(u))).T_{\alpha}(u)=b_{\alpha}(u)\cdot\nabla_{\mathbb{S}}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u)=b_{\alpha}(u)\cdot\pi_{H}(\pi_{u^{\perp}}(\nabla\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u))).

Since now b(u)Hb(u)\in H and since πH\pi_{H} and πu\pi_{u^{\perp}} are self-adjoint, as every orthogonal projection, we get

Tα(u)=πu(bα(u))Γ𝐊,ε𝕊,n(u)=bα(u)Γ𝐊,ε𝕊,n(u)(bα(u)u)(uΓ𝐊,ε𝕊,n(u)).T_{\alpha}(u)=\pi_{u^{\perp}}(b_{\alpha}(u))\cdot\nabla\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u)=b_{\alpha}(u)\cdot\nabla\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u)-(b_{\alpha}(u)\cdot u)(u\cdot\nabla\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u)).

But bα(u)Γ𝐊,ε𝕊,n(u)=αXΓ𝐊,ε𝕊,n(u)12ΔΓ𝐊,ε𝕊,n(u)b_{\alpha}(u)\cdot\nabla\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u)={\mathcal{L}}^{X}_{\alpha}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u)-\frac{1}{2}\Delta\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u) is uniformly bounded by point (i) and since ΔΓ𝐊,ε𝕊,n(u)\Delta\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u) is bounded on 𝕊{\mathbb{S}}. Next, uΓ𝐊,ε𝕊,n(u)u\cdot\nabla\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}(u) is smooth and thus bounded on 𝕊{\mathbb{S}}. Finally,

bα(u)u=θN1i,jN(uiuj)uiuiuj2+α=θ2N1i,jNuiuj2uiuj2+αb_{\alpha}(u)\cdot u=-\frac{\theta}{N}\sum_{1\leq i,j\leq N}\frac{(u^{i}-u^{j})\cdot u^{i}}{\|u^{i}-u^{j}\|^{2}+\alpha}=-\frac{\theta}{2N}\sum_{1\leq i,j\leq N}\frac{\|u^{i}-u^{j}\|^{2}}{\|u^{i}-u^{j}\|^{2}+\alpha}

is also uniformly bounded. ∎

Remark 6.3.

We have proved in Step 2 that for each m>0m>0, gmCc((2)N)g_{m}\in C^{\infty}_{c}(({\mathbb{R}}^{2})^{N}) satisfies gm=1g_{m}=1 on B(0,m)B(0,m) and supα(0,1]supx(2)N|αXgm(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}g_{m}(x)|<\infty.

7. A Girsanov theorem for the Keller-Segel particle system.

In this section, we prove a rigorous version of the intuitive argument presented in Subsection 3.4.

For x(2)Nx\in({\mathbb{R}}^{2})^{N}, all K[[1,N]]K\subset[\![1,N]\!], we denote by x|K=(xi)iKx|_{K}=(x^{i})_{i\in K}. For 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} a partition of [[1,N]][\![1,N]\!], for y1(2)|K1|,,y(2)|K|y_{1}\in({\mathbb{R}}^{2})^{|K_{1}|},\dots,y_{\ell}\in({\mathbb{R}}^{2})^{|K_{\ell}|}, we abusively denote by (yp)p[[1,]](y_{p})_{p\in[\![1,\ell]\!]} the element yy of (2)N({\mathbb{R}}^{2})^{N} such that for all i[[1,]]i\in[\![1,\ell]\!], y|Ki=yiy|_{K_{i}}=y_{i}.

We adopt the convention that for any θ>0\theta>0, a QKS(θ,1)QKS(\theta,1)-process is a 22-dimensional Brownian motion. This is natural in view of (1).

Proposition 7.1.

Let N2N\geq 2, θ>0\theta>0 such that N>θN>\theta and set k0=2N/θk_{0}=\lceil 2N/\theta\rceil. Fix some partition 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} of [[1,N]][\![1,N]\!] with 2\ell\geq 2. Consider the state spaces 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} and, for each p[[1,]]p\in[\![1,\ell]\!],

𝒴p={y(2)|Kp|:K[[1,|Kp|]] with |K|k0i,j=1|Kp|yiyj2>0}.{\mathcal{Y}}_{p}=\Big{\{}y\in({\mathbb{R}}^{2})^{|K_{p}|}:\forall K\subset[\![1,|K_{p}|]\!]\hbox{ with $|K|\geq k_{0}$, }\sum_{i,j=1}^{|K_{p}|}||y^{i}-y^{j}||^{2}>0\Big{\}}.

Consider

\bullet 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳){\mathbb{X}}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}^{X}_{x})_{x\in{\mathcal{X}_{\triangle}}}) a QKS(θ,N)QKS(\theta,N)-process,

\bullet For all p[[1,]]p\in[\![1,\ell]\!], 𝕐p=(Ωp,p,(Yp,t)t0,(yp)y𝒴p){\mathbb{Y}}^{p}=(\Omega^{p},{\mathcal{M}}^{p},(Y_{p,t})_{t\geq 0},({\mathbb{P}}^{p}_{y})_{y\in{\mathcal{Y}}^{p}_{\triangle}}) a QKS(θ|Kp|/N,|Kp|)QKS(\theta|K_{p}|/N,|K_{p}|)-process.

We set ΩY=p=1Ωp\Omega^{Y}=\prod_{p=1}^{\ell}\Omega^{p} and Yt=(Yp,t)p[[1,]]Y_{t}=(Y_{p,t})_{p\in[\![1,\ell]\!]}, with the convention that Yt=Y_{t}=\triangle as soon as Yp,t=Y_{p,t}=\triangle for some p[[1,]]p\in[\![1,\ell]\!]. We also introduce Y=σ(Yt:t0){\mathcal{M}}^{Y}=\sigma(Y_{t}:t\geq 0), as well as yY=p=1ypp{\mathbb{P}}^{Y}_{y}=\otimes_{p=1}^{\ell}{\mathbb{P}}^{p}_{y_{p}} for all y=(yp)p[[1,]](2)Ny=(y_{p})_{p\in[\![1,\ell]\!]}\in({\mathbb{R}}^{2})^{N}.

We fix ε(0,1]\varepsilon\in(0,1], recall that

G𝐊,ε={x𝒳:min1pqminiKp,jKqxixj2>ε}B(0,1ε),G_{{\mathbf{K}},\varepsilon}=\Big{\{}x\in{\mathcal{X}}:\min_{1\leq p\neq q\leq\ell}\;\;\min_{i\in K_{p},j\in K_{q}}\|x^{i}-x^{j}\|^{2}>\varepsilon\Big{\}}\cap B\Big{(}0,\frac{1}{\varepsilon}\Big{)},

and set

τ𝐊,ε={t0:XtG𝐊,ε}andτ~𝐊,ε={t0:YtG𝐊,ε}.\displaystyle\tau_{{\mathbf{K}},\varepsilon}=\big{\{}t\geq 0:X_{t}\notin G_{{\mathbf{K}},\varepsilon}\}\quad\hbox{and}\quad\tilde{\tau}_{{\mathbf{K}},\varepsilon}=\big{\{}t\geq 0:Y_{t}\notin G_{{\mathbf{K}},\varepsilon}\}.

Fix T>0T>0. Quasi-everywhere in G𝐊,εG_{{\mathbf{K}},\varepsilon}, there is a probability measure xT,ε,𝐊\mathbb{Q}_{x}^{T,\varepsilon,{\mathbf{K}}} on (ΩX,X)(\Omega^{X},{\mathcal{M}}^{X}), equivalent to xX{\mathbb{P}}_{x}^{X}, such that the law of the process (XtTτ𝐊,ε)t0(X_{t\land T\land\tau_{{\mathbf{K}},\varepsilon}})_{t\geq 0} under xT,ε,𝐊\mathbb{Q}_{x}^{T,\varepsilon,{\mathbf{K}}} is the same as that of (YtTτ~𝐊,ε)t0(Y_{t\land T\land\tilde{\tau}_{{\mathbf{K}},\varepsilon}})_{t\geq 0} on (ΩY,Y)(\Omega^{Y},{\mathcal{M}}^{Y}) under xY{\mathbb{P}}^{Y}_{x}.

Furthermore, the Radon-Nikodym density dxT,ε,𝐊dxX\frac{{\rm d}\mathbb{Q}^{T,\varepsilon,{\mathbf{K}}}_{x}}{{\rm d}{\mathbb{P}}^{X}_{x}} is Tτ𝐊,εX{\mathcal{M}}^{X}_{T\land\tau_{{\mathbf{K}},\varepsilon}}-measurable, where as usual tX=σ(Xs,st){\mathcal{M}}^{X}_{t}={\sigma(X_{s},s\leq t)}, and there is a deterministic constant CT,ε,𝐊>0C_{T,\varepsilon,{\mathbf{K}}}>0 such that quasi-everywhere in G𝐊,εG_{{\mathbf{K}},\varepsilon},

CT,ε,𝐊1dxT,ε,𝐊dxXCT,ε,𝐊.C_{T,\varepsilon,{\mathbf{K}}}^{-1}\leq\frac{d\mathbb{Q}^{T,\varepsilon,{\mathbf{K}}}_{x}}{d{\mathbb{P}}^{X}_{x}}\leq C_{T,\varepsilon,{\mathbf{K}}}.

The quasi-everywhere notion refers to the process 𝕏{\mathbb{X}}. Let us mention that for ζ\zeta the life-time of 𝕏{\mathbb{X}}, we have τ𝐊,ε[0,ζ]\tau_{{\mathbf{K}},\varepsilon}\in[0,\zeta] when ζ<\zeta<\infty because G𝐊,ε\triangle\notin G_{{\mathbf{K}},\varepsilon}. Although this is not clear at this point of the paper, the event {τ𝐊,ε=ζ}\{\tau_{{\mathbf{K}},\varepsilon}=\zeta\} has a positive probability if maxp=1,,|Kp|k0\max_{p=1,\dots,\ell}|K_{p}|\geq k_{0}.

Proof.

We only consider the case where =2\ell=2. The general case is heavier in terms of notation but contains no additional difficulty. We fix 𝐊=(K1,K2){\mathbf{K}}=(K_{1},K_{2}) a non-trivial partition of [[1,N]][\![1,N]\!]. The main idea is to apply Lemma B.7 to 𝕏{\mathbb{X}} with the function

(1) ϱ(x)=exp(u(x)),whereu(x)=θNiK1,jK2log(xixj).\varrho(x)=\exp(u(x)),\quad\hbox{where}\quad u(x)=\frac{\theta}{N}\sum_{i\in K_{1},j\in K_{2}}\log(\|x^{i}-x^{j}\|).

Unfortunately, this is not licit because uXu\notin{\mathcal{F}}^{X}.

Step 1. Set 𝕐=(ΩY,Y,(Yt)t0,(yY)y(𝒴1×𝒴2){}){\mathbb{Y}}=(\Omega^{Y},{\mathcal{M}}^{Y},(Y_{t})_{t\geq 0},({\mathbb{P}}^{Y}_{y})_{y\in({\mathcal{Y}}_{1}\times{\mathcal{Y}}_{2})\cup\{\triangle\}}) and fix ε(0,1]\varepsilon\in(0,1] and n1n\geq 1. We first compute the Dirichlet space of 𝕐{\mathbb{Y}} killed when it gets outside of G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n}, recall Lemma 6.1. Consider the measures

μ1(dy)=i,jK1,ijyiyjθ/Ndyandμ2(dy)=i,jK2,ijyiyjθ/Ndy\mu_{1}({\rm d}y)=\prod_{i,j\in K_{1},i\neq j}||y^{i}-y^{j}||^{-\theta/N}{\rm d}y\quad\hbox{and}\quad\mu_{2}({\rm d}y)=\prod_{i,j\in K_{2},i\neq j}||y^{i}-y^{j}||^{-\theta/N}{\rm d}y

on (2)|K1|({\mathbb{R}}^{2})^{|K_{1}|} and (2)|K2|({\mathbb{R}}^{2})^{|K_{2}|}, with μi(dy)=dy\mu_{i}({\rm d}y)={\rm d}y if |Ki|=1|K_{i}|=1. Recall that μ(dx)=𝐦(x)dx\mu({\rm d}x)={\mathbf{m}}(x){\rm d}x, see (4) and that by definition, see (1), ϱ(x)=iK1,jK2xixjθ/N\varrho(x)=\prod_{i\in K_{1},j\in K_{2}}\|x^{i}-x^{j}\|^{\theta/N}: we deduce that

μ1μ2=ϱ2μ.\mu_{1}\otimes\mu_{2}=\varrho^{2}\mu.

By Proposition 4.1, for p=1,2p=1,2, 𝕐p{\mathbb{Y}}^{p} is a 𝒴p{\mathcal{Y}}^{p}_{\triangle}-valued μp\mu_{p}-symmetric (since (θ|Kp|/N)/|Kp|=θ/N(\theta|K_{p}|/N)/|K_{p}|=\theta/N) diffusion with regular Dirichlet space (p,p)({\mathcal{E}}_{p},{\mathcal{F}}_{p}) with core Cc(𝒴p)C^{\infty}_{c}({\mathcal{Y}}_{p}) and, for φCc(𝒴p)\varphi\in C^{\infty}_{c}({\mathcal{Y}}_{p}), p(φ,φ)=12(2)|Kp|φ2dμp{\mathcal{E}}_{p}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{|K_{p}|}}||\nabla\varphi||^{2}{\rm d}\mu_{p}. This also holds true if e.g. |K1|=1|K_{1}|=1, see [11, Example 4.2.1 page 167], since then μ1\mu_{1} is nothing but the Lebesgue measure on 2{\mathbb{R}}^{2}. Since now μ1μ2=ϱ2μ\mu_{1}\otimes\mu_{2}=\varrho^{2}\mu, by Lemma B.5, 𝕐{\mathbb{Y}} is a ϱ2μ\varrho^{2}\mu-symmetric 𝒳{\mathcal{X}}_{\triangle}-valued diffusion with regular Dirichlet space (Y,Y)({\mathcal{E}}^{Y},{\mathcal{F}}^{Y}) on L2(𝒴1×𝒴2,ϱ2dμ)L^{2}({\mathcal{Y}}_{1}\times{\mathcal{Y}}_{2},\varrho^{2}{\rm d}\mu) with core Cc(𝒴1×𝒴2)C^{\infty}_{c}({\mathcal{Y}}_{1}\times{\mathcal{Y}}_{2}) and, for φCc(𝒴1×𝒴2)\varphi\in C^{\infty}_{c}({\mathcal{Y}}_{1}\times{\mathcal{Y}}_{2}),

Y(φ,φ)=\displaystyle{\mathcal{E}}^{Y}(\varphi,\varphi)= (2)|K1|2(φ(y,),φ(y,))μ1(dy)+(2)|K2|1(φ(,z),φ(,z))μ2(dz)=12(2)Nφ2ϱ2dμ.\displaystyle\int_{({\mathbb{R}}^{2})^{|K_{1}|}}\hskip-8.5359pt{\mathcal{E}}_{2}(\varphi(y,\cdot),\varphi(y,\cdot))\mu_{1}({\rm d}y)+\int_{({\mathbb{R}}^{2})^{|K_{2}|}}\hskip-8.5359pt{\mathcal{E}}_{1}(\varphi(\cdot,z),\varphi(\cdot,z))\mu_{2}({\rm d}z)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\hskip-8.5359pt\|\nabla\varphi\|^{2}\varrho^{2}{\rm d}\mu.

Finally, we apply Lemma B.6 to 𝕐{\mathbb{Y}} with the open set G𝐊,εn𝒳𝒴1×𝒴2G_{{\mathbf{K}},\varepsilon}^{n}\subset{\mathcal{X}}\subset{\mathcal{Y}}_{1}\times{\mathcal{Y}}_{2}, to find that the resulting killed process

𝕐n,ε=(ΩY,Y,(Ytn,ε)t0,(yY)yG𝐊,εn{}){\mathbb{Y}}^{n,\varepsilon}=\Big{(}\Omega^{Y},{\mathcal{M}}^{Y},(Y_{t}^{n,\varepsilon})_{t\geq 0},({\mathbb{P}}^{Y}_{y})_{y\in G_{{\mathbf{K}},\varepsilon}^{n}\cup\{\triangle\}}\Big{)}

is a ϱ2μ|G𝐊,εn\varrho^{2}\mu|_{G_{{\mathbf{K}},\varepsilon}^{n}}-symmetric G𝐊,εn{}G_{{\mathbf{K}},\varepsilon}^{n}\cup\{\triangle\}-valued diffusion with regular Dirichlet space (Y,n,ε,Y,n,ε)({{\mathcal{E}}}^{Y,n,\varepsilon},{{\mathcal{F}}}^{Y,n,\varepsilon}) with core Cc(G𝐊,εn)C_{c}^{\infty}(G_{{\mathbf{K}},\varepsilon}^{n}) such that for all φCc(G𝐊,εn)\varphi\in C_{c}^{\infty}(G_{{\mathbf{K}},\varepsilon}^{n}),

Y,n,ε(φ,φ)=12G𝐊,εnφ2ϱ2dμ.{{\mathcal{E}}}^{Y,n,\varepsilon}(\varphi,\varphi)=\frac{1}{2}\int_{G_{{\mathbf{K}},\varepsilon}^{n}}||\nabla\varphi||^{2}\varrho^{2}{\rm d}\mu.

Step 2. We now fix ε(0,1]\varepsilon\in(0,1] and introduce, for each n1n\geq 1, un,ε(x)=u(x)Γ𝐊,εn(x)u_{n,\varepsilon}(x)=u(x)\Gamma_{{\mathbf{K}},\varepsilon}^{n}(x), recall (1) and Lemma 6.1, and ϱn,ε=exp(un,ε)\varrho_{n,\varepsilon}=\exp(u_{n,\varepsilon}). We check here that the functions un,εu_{n,\varepsilon} and ϱn,ε\varrho_{n,\varepsilon} satisfy the assumptions of Lemma B.7 (to be applied to 𝕏{\mathbb{X}}), that 𝒜X[ϱn,ε1]=Xϱn,ε{\mathcal{A}}^{X}[\varrho_{n,\varepsilon}-1]={\mathcal{L}}^{X}\varrho_{n,\varepsilon} and that

(2) supn1supx𝒳|un,ε(x)|<andsupn1supxG𝐊,εn|Xϱn,ε(x)|<.\sup_{n\geq 1}\sup_{x\in{\mathcal{X}}}|u_{n,\varepsilon}(x)|<\infty\qquad\hbox{and}\qquad\sup_{n\geq 1}\sup_{x\in G_{{\mathbf{K}},\varepsilon}^{n}}|{\mathcal{L}}^{X}\varrho_{n,\varepsilon}(x)|<\infty.

First, un,εXu_{n,\varepsilon}\in{\mathcal{F}}^{X} because un,εCc(𝒳)u_{n,\varepsilon}\in C^{\infty}_{c}({\mathcal{X}}), and |un,ε||u_{n,\varepsilon}| is bounded, uniformly in n1n\geq 1, because Γ𝐊,εn\Gamma_{{\mathbf{K}},\varepsilon}^{n} is bounded by 11 and vanishes outside G𝐊,ηG_{{\mathbf{K}},\eta} (see Lemma 6.1), while uu is smooth on G𝐊,ηG_{{\mathbf{K}},\eta}. To show that 𝒜X[ϱn,ε1]=Xϱn,ε{\mathcal{A}}^{X}[\varrho_{n,\varepsilon}-1]={\mathcal{L}}^{X}\varrho_{n,\varepsilon}, it suffices by Remark 4.3 to verify that ϱn,ε1Cc(𝒳)\varrho_{n,\varepsilon}-1\in C^{\infty}_{c}({\mathcal{X}}), which is clear, and that supα(0,1]supx(2)N|αXϱn,ε(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}_{\alpha}^{X}\varrho_{n,\varepsilon}(x)|<\infty. We have

αXϱn,ε(x)=eun,ε(x)αXun,ε(x)+12eun,ε(x)un,ε(x)2.{\mathcal{L}}^{X}_{\alpha}\varrho_{n,\varepsilon}(x)=e^{u_{n,\varepsilon}(x)}{\mathcal{L}}^{X}_{\alpha}u_{n,\varepsilon}(x)+\frac{1}{2}e^{u_{n,\varepsilon}(x)}\|\nabla u_{n,\varepsilon}(x)\|^{2}.

Since un,εCc((2)N)u_{n,\varepsilon}\in C_{c}^{\infty}(({\mathbb{R}}^{2})^{N}), the only difficulty is to check that supα(0,1]supx(2)N|αXun,ε(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in({\mathbb{R}}^{2})^{N}}|{\mathcal{L}}^{X}_{\alpha}u_{n,\varepsilon}(x)|<\infty. By (3),

αXun,ε(x)=Γ𝐊,εn(x)αXu(x)+u(x)αXΓ𝐊,εn(x)+Γ𝐊,εn(x)u(x).{\mathcal{L}}^{X}_{\alpha}u_{n,\varepsilon}(x)=\Gamma_{{\mathbf{K}},\varepsilon}^{n}(x){\mathcal{L}}^{X}_{\alpha}u(x)+u(x){\mathcal{L}}^{X}_{\alpha}\Gamma_{{\mathbf{K}},\varepsilon}^{n}(x)+\nabla\Gamma_{{\mathbf{K}},\varepsilon}^{n}(x)\cdot\nabla u(x).

Again, the only difficulty consists of the first term, because αXΓ𝐊,εn{\mathcal{L}}^{X}_{\alpha}\Gamma_{{\mathbf{K}},\varepsilon}^{n} is uniformly bounded by Lemma 6.1 and vanishes outside G𝐊,ηG_{{\mathbf{K}},\eta}, while uu is smooth on G𝐊,ηG_{{\mathbf{K}},\eta}. Since Supp Γ𝐊,εnG𝐊,η\Gamma_{{\mathbf{K}},\varepsilon}^{n}\subset G_{{\mathbf{K}},\eta}, we are reduced to show that supα(0,1]supxG𝐊,η|αXu(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in G_{{\mathbf{K}},\eta}}|{\mathcal{L}}^{X}_{\alpha}u(x)|<\infty. But

αXu=12ΔuθNSα,whereSα(x)=1i,jNxixjxixj2+αxiu(x),{\mathcal{L}}^{X}_{\alpha}u=\frac{1}{2}\Delta u-\frac{\theta}{N}S_{\alpha},\quad\hbox{where}\quad S_{\alpha}(x)=\sum_{1\leq i,j\leq N}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot\nabla_{x^{i}}{u}(x),

and we only have to verify that supα(0,1]supxG𝐊,η|Sα(x)|<\sup_{\alpha\in(0,1]}\sup_{x\in G_{{\mathbf{K}},\eta}}|S_{\alpha}(x)|<\infty.

For kK1k\in K_{1} and K2\ell\in K_{2}, we have

xku(x)=jK2θNxkxjxkxj2andxu(x)=iK1θNxxixxi2.\displaystyle\nabla_{x^{k}}{u}(x)=\sum_{j\in K_{2}}\frac{\theta}{N}\frac{x^{k}-x^{j}}{\|x^{k}-x^{j}\|^{2}}\quad\hbox{and}\quad\nabla_{x^{\ell}}{u}(x)=\sum_{i\in K_{1}}\frac{\theta}{N}\frac{x^{\ell}-x^{i}}{\|x^{\ell}-x^{i}\|^{2}}.

Hence Sα=S1,α+S2,α+S3,α+S4,αS_{\alpha}=S_{1,\alpha}+S_{2,\alpha}+S_{3,\alpha}+S_{4,\alpha}, where

S1,α(x)=\displaystyle S_{1,\alpha}(x)= θNi,jK1xixjxixj2+αkK2xixkxixk2,\displaystyle\frac{\theta}{N}\sum_{i,j\in K_{1}}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot\sum_{k\in K_{2}}\frac{x^{i}-x^{k}}{\|x^{i}-x^{k}\|^{2}},
S2,α(x)=\displaystyle S_{2,\alpha}(x)= θNiK2,jK1xixjxixj2+αkK1xixkxixk2,\displaystyle\frac{\theta}{N}\sum_{i\in K_{2},j\in K_{1}}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\cdot\sum_{k\in K_{1}}\frac{x^{i}-x^{k}}{\|x^{i}-x^{k}\|^{2}},

and S3,αS_{3,\alpha} (resp. S4,αS_{4,\alpha}) is defined as S1,αS_{1,\alpha} (resp. S2,αS_{2,\alpha}) exchanging the roles of K1K_{1} and K2K_{2}. First, S2,αS_{2,\alpha} (and S4,αS_{4,\alpha}) is obviously uniformly bounded on G𝐊,ηG_{{\mathbf{K}},\eta}. Next, by symmetry,

S1,α(x)=θ2Ni,jK1xixjxixj2+αkK2(xixkxixk2xjxkxjxk2).S_{1,\alpha}(x)=\frac{\theta}{2N}\sum_{i,j\in K_{1}}\frac{x^{i}-x^{j}}{\|x^{i}-x^{j}\|^{2}+\alpha}\sum_{k\in K_{2}}\Big{(}\frac{x^{i}-x^{k}}{\|x^{i}-x^{k}\|^{2}}-\frac{x^{j}-x^{k}}{\|x^{j}-x^{k}\|^{2}}\Big{)}.

Moreover, there is Cη>0C_{\eta}>0 such that for all xG𝐊,ηx\in G_{{\mathbf{K}},\eta}, all i,jK1i,j\in K_{1} such that iji\neq j, all kK2k\in K_{2},

xixkxixk2xjxkxjxk2Cηxixj,\displaystyle\Big{\|}\frac{x^{i}-x^{k}}{\|x^{i}-x^{k}\|^{2}}-\frac{x^{j}-x^{k}}{\|x^{j}-x^{k}\|^{2}}\Big{\|}\leq C_{\eta}\|x^{i}-x^{j}\|,

so that S1,αS_{1,\alpha} (and S3,αS_{3,\alpha}) is bounded on G𝐊,ηG_{{\mathbf{K}},\eta}, uniformly in α(0,1]\alpha\in(0,1], as desired. Finally, the above computations, together with the facts that Γ𝐊,εn=1\Gamma_{{\mathbf{K}},\varepsilon}^{n}=1 on G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n}, also show that for xG𝐊,εnx\in G_{{\mathbf{K}},\varepsilon}^{n},

Xϱn,ε(x)=eu(x)(12Δu(x)θNSα(x))+12eu(x)u(x)2,{\mathcal{L}}^{X}\varrho_{n,\varepsilon}(x)=e^{u(x)}\Big{(}\frac{1}{2}\Delta u(x)-\frac{\theta}{N}S_{\alpha}(x)\Big{)}+\frac{1}{2}e^{u(x)}||\nabla u(x)||^{2},

which is bounded on G𝐊,ηG_{{\mathbf{K}},\eta}. Since G𝐊,εnG𝐊,ηG_{{\mathbf{K}},\varepsilon}^{n}\subset G_{{\mathbf{K}},\eta}, this implies that supn1supxG𝐊,εn|Xϱn,ε(x)|\sup_{n\geq 1}\sup_{x\in G_{{\mathbf{K}},\varepsilon}^{n}}|{\mathcal{L}}^{X}\varrho_{n,\varepsilon}(x)| and completes the step.

Step 3. We apply Lemma B.7 to the process 𝕏{\mathbb{X}} with un,εu_{n,\varepsilon} and ϱn,ε\varrho_{n,\varepsilon} defined in Step 2. Recalling that 𝒜X[ϱn,ε1]=Xϱn,ε{\mathcal{A}}^{X}[\varrho_{n,\varepsilon}-1]={\mathcal{L}}^{X}\varrho_{n,\varepsilon} and using the conventions ϱn,ε()=1\varrho_{n,\varepsilon}(\triangle)=1 and Xϱn,ε()=0{\mathcal{L}}^{X}\varrho_{n,\varepsilon}(\triangle)=0, we set

(3) Ltn,ε=ϱn,ε(Xt)ϱn,ε(X0)exp(0tXϱn,ε(Xs)ϱn,ε(Xs)ds).L^{n,\varepsilon}_{t}=\frac{\varrho_{n,\varepsilon}(X_{t})}{\varrho_{n,\varepsilon}(X_{0})}\exp\Big{(}-\int_{0}^{t}\frac{{\mathcal{L}}^{X}\varrho_{n,\varepsilon}(X_{s})}{\varrho_{n,\varepsilon}(X_{s})}{\rm d}s\Big{)}.

Set tX=σ({Xs,st}){\mathcal{M}}^{X}_{t}=\sigma(\{X_{s},s\leq t\}). By Lemma B.7, there is a family of probability measures (xn,ε)x𝒳{}({\mathbb{Q}}^{n,\varepsilon}_{x})_{x\in{\mathcal{X}}\cup\{\triangle\}} such that

xn,ε=Ltn,εxXontX{\mathbb{Q}}^{n,\varepsilon}_{x}=L_{t}^{n,\varepsilon}\cdot{\mathbb{P}}_{x}^{X}\quad\hbox{on}\quad{\mathcal{M}}^{X}_{t}

for all t0t\geq 0 and quasi-everywhere in 𝒳{}{\mathcal{X}}\cup\{\triangle\}, and such that

𝕏n,ε=(ΩX,X,(Xt)t0,(xn,ε)x𝒳){\mathbb{X}}^{n,\varepsilon}=\Big{(}\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{Q}}^{n,\varepsilon}_{x})_{x\in{\mathcal{X}}_{\triangle}}\Big{)}

is a ϱn,ε2μ\varrho_{n,\varepsilon}^{2}\mu-symmetric 𝒳{}{\mathcal{X}}\cup\{\triangle\}-valued diffusion with regular Dirichlet space (n,ε,n,ε)({\mathcal{E}}^{n,\varepsilon},{\mathcal{F}}^{n,\varepsilon}) with core Cc(𝒳)C_{c}^{\infty}({\mathcal{X}}) such that for all φCc(𝒳)\varphi\in C_{c}^{\infty}({\mathcal{X}}),

n,ε(φ,φ)=12(2)Nφ2ϱn,ε2dμ.{\mathcal{E}}^{n,\varepsilon}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}||\nabla\varphi||^{2}\varrho_{n,\varepsilon}^{2}{\rm d}\mu.

Next, we apply Lemma B.6 to 𝕏n,ε{\mathbb{X}}^{n,\varepsilon} with the open set G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n}: the resulting killed process

𝕏,n,ε=(ΩX,X,(Xt,n,ε)t0,(xn,ε)xG𝐊,εn{}){\mathbb{X}}^{*,n,\varepsilon}=\Big{(}\Omega^{X},{\mathcal{M}}^{X},(X_{t}^{*,n,\varepsilon})_{t\geq 0},({\mathbb{Q}}^{n,\varepsilon}_{x})_{x\in G_{{\mathbf{K}},\varepsilon}^{n}\cup\{\triangle\}}\Big{)}

is a ϱn,ε2μ|G𝐊,εn\varrho_{n,\varepsilon}^{2}\mu|_{G_{{\mathbf{K}},\varepsilon}^{n}}-symmetric G𝐊,εn{}G_{{\mathbf{K}},\varepsilon}^{n}\cup\{\triangle\}-valued diffusion with regular Dirichlet space (,n,ε,,n,ε)({{\mathcal{E}}}^{*,n,\varepsilon},{{\mathcal{F}}}^{*,n,\varepsilon}) with core Cc(G𝐊,εn)C_{c}^{\infty}(G_{{\mathbf{K}},\varepsilon}^{n}) such that for all φCc(G𝐊,εn)\varphi\in C_{c}^{\infty}(G_{{\mathbf{K}},\varepsilon}^{n}),

,n,ε(φ,φ)=12G𝐊,εnφ2ϱn,ε2dμ.{{\mathcal{E}}}^{*,n,\varepsilon}(\varphi,\varphi)=\frac{1}{2}\int_{G_{{\mathbf{K}},\varepsilon}^{n}}||\nabla\varphi||^{2}\varrho_{n,\varepsilon}^{2}{\rm d}\mu.

Comparing this Dirichlet space with the one found in Step 1, using that ϱn,ε=ϱ\varrho_{n,\varepsilon}=\varrho on G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n} and a uniqueness argument, see [11, Theorem 4.2.8 p 167], we conclude that quasi-everywhere in G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n}, the law of X,n,εX^{*,n,\varepsilon} under xn,ε{\mathbb{Q}}^{n,\varepsilon}_{x} equals the law of Yn,εY^{n,\varepsilon} under xY{\mathbb{P}}^{Y}_{x}.

Step 4. We fix T>0T>0 and ε(0,1]\varepsilon\in(0,1] and complete the proof. Since xn,ε=LTn,εxX{\mathbb{Q}}^{n,\varepsilon}_{x}=L^{n,\varepsilon}_{T}\cdot{\mathbb{P}}_{x}^{X} on TX{\mathcal{M}}^{X}_{T}, we know from Step 3 that for all n1n\geq 1, quasi-everywhere in G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n}, for all continuous bounded Φ:C([0,T],𝒳)\Phi:C([0,T],{\mathcal{X}}_{\triangle})\to{\mathbb{R}}, (observe that G¯𝐊,εn𝒳𝒳\bar{G}_{{\mathbf{K}},\varepsilon}^{n}\subset{\mathcal{X}}\subset{\mathcal{X}}_{\triangle})

𝔼xX[Φ(Xτ𝐊,n,εT)LTn,ε]=𝔼xY[Φ(Yτ~𝐊,n,εT)],{\mathbb{E}}_{x}^{X}[\Phi(X_{\cdot\land\tau_{{\mathbf{K}},n,\varepsilon}\land T})L^{n,\varepsilon}_{T}]={\mathbb{E}}_{x}^{Y}[\Phi(Y_{\cdot\land\tilde{\tau}_{{\mathbf{K}},n,\varepsilon}\land T})],

where τ𝐊,n,ε=inf{t>0:XtG𝐊,εn}τ𝐊,ε\tau_{{\mathbf{K}},n,\varepsilon}=\inf\{t>0:X_{t}\notin G_{{\mathbf{K}},\varepsilon}^{n}\}\land\tau_{{\mathbf{K}},\varepsilon} and τ~𝐊,n,ε=inf{t>0:YtG𝐊,εn}τ~𝐊,ε\tilde{\tau}_{{\mathbf{K}},n,\varepsilon}=\inf\{t>0:Y_{t}\notin G_{{\mathbf{K}},\varepsilon}^{n}\}\land\tilde{\tau}_{{\mathbf{K}},\varepsilon}. Since (Ltn,ε)t0(L^{n,\varepsilon}_{t})_{t\geq 0} is a xX{\mathbb{P}}_{x}^{X}-martingale by Lemma B.7, we deduce that quasi-everywhere in G𝐊,εnG_{{\mathbf{K}},\varepsilon}^{n},

(4) 𝔼xX[Φ(Xτ𝐊,n,εT)Lτ𝐊,n,εTn,ε]=𝔼xY[Φ(Yτ~𝐊,n,εT)].{\mathbb{E}}_{x}^{X}[\Phi(X_{\cdot\land\tau_{{\mathbf{K}},n,\varepsilon}\land T})L^{n,\varepsilon}_{\tau_{{\mathbf{K}},n,\varepsilon}\land T}]={\mathbb{E}}_{x}^{Y}[\Phi(Y_{\cdot\land\tilde{\tau}_{{\mathbf{K}},n,\varepsilon}\land T})].

Recall that G𝐊,εn1G𝐊,εnG_{{\mathbf{K}},\varepsilon}\subset\cup_{n\geq 1}G_{{\mathbf{K}},\varepsilon}^{n}, see Lemma 6.1. Hence limnτ𝐊,n,ε=τ𝐊,ε\lim_{n}\tau_{{\mathbf{K}},n,\varepsilon}=\tau_{{\mathbf{K}},\varepsilon}, limnτ~𝐊,n,ε=τ~𝐊,ε\lim_{n}\tilde{\tau}_{{\mathbf{K}},n,\varepsilon}=\tilde{\tau}_{{\mathbf{K}},\varepsilon}, and for each xG𝐊,εx\in G_{{\mathbf{K}},\varepsilon}, there is nx1n_{x}\geq 1 such that xG𝐊,εnx\in G_{{\mathbf{K}},\varepsilon}^{n} for all nnxn\geq n_{x}. We deduce from (4) that quasi-everywhere in G𝐊,εG_{{\mathbf{K}},\varepsilon}, the process (Lτ𝐊,n,εTn,ε)nnx(L^{n,\varepsilon}_{\tau_{{\mathbf{K}},n,\varepsilon}\land T})_{n\geq n_{x}} is a (τ𝐊,n,εTX)nnx({\mathcal{M}}_{\tau_{{\mathbf{K}},n,\varepsilon}\land T}^{X})_{n\geq n_{x}}-martingale under xX{\mathbb{P}}_{x}^{X}. Moreover, recalling the expression (3) of Ln,εL^{n,\varepsilon}, that ϱn,ε=exp(un,ε)\varrho_{n,\varepsilon}=\exp(u_{n,\varepsilon}) and the bound (2), we conclude that there is a constant CT,ε,𝐊>0C_{T,\varepsilon,{\mathbf{K}}}>0 such that quasi-everywhere in G𝐊,εG_{{\mathbf{K}},\varepsilon},

xX-a.s., for all nnx,CT,ε,𝐊1Lτ𝐊,n,εTn,εCT,ε,𝐊.\hbox{${\mathbb{P}}^{X}_{x}$-a.s., for all $n\geq n_{x}$,}\quad C_{T,\varepsilon,{\mathbf{K}}}^{-1}\leq L^{n,\varepsilon}_{\tau_{{\mathbf{K}},n,\varepsilon}\land T}\leq C_{T,\varepsilon,{\mathbf{K}}}.

Hence the martingale (Lτ𝐊,n,εTn,ε)nnx(L^{n,\varepsilon}_{\tau_{{\mathbf{K}},n,\varepsilon}\land T})_{n\geq n_{x}} is closed by some τ𝐊,εT{\mathcal{M}}_{\tau_{{\mathbf{K}},\varepsilon}\land T}-measurable random variable JT,ε,𝐊J_{T,\varepsilon,{\mathbf{K}}} that satisfies CT,ε,𝐊1JT,ε,𝐊CT,ε,𝐊C_{T,\varepsilon,{\mathbf{K}}}^{-1}\leq J_{T,\varepsilon,{\mathbf{K}}}\leq C_{T,\varepsilon,{\mathbf{K}}}, and (4) implies that for all nnxn\geq n_{x},

𝔼xX[Φ(Xτ𝐊,n,εT)JT,ε,𝐊]=𝔼xY[Φ(Yτ~𝐊,n,εT)].{\mathbb{E}}_{x}^{X}[\Phi(X_{\cdot\land\tau_{{\mathbf{K}},n,\varepsilon}\land T})J_{T,\varepsilon,{\mathbf{K}}}]={\mathbb{E}}_{x}^{Y}[\Phi(Y_{\cdot\land\tilde{\tau}_{{\mathbf{K}},n,\varepsilon}\land T})].

Letting nn\to\infty, we find that quasi-everywhere in G𝐊,εG_{{\mathbf{K}},\varepsilon}, for ΦCb(C([0,T],𝒳),)\Phi\in C_{b}(C([0,T],{\mathcal{X}}_{\triangle}),{\mathbb{R}}),

𝔼xX[Φ(Xτ𝐊,εT)JT,ε,𝐊]=𝔼xY[Φ(Yτ~𝐊,εT)].{\mathbb{E}}_{x}^{X}[\Phi(X_{\cdot\land\tau_{{\mathbf{K}},\varepsilon}\land T})J_{T,\varepsilon,{\mathbf{K}}}]={\mathbb{E}}_{x}^{Y}[\Phi(Y_{\cdot\land\tilde{\tau}_{{\mathbf{K}},\varepsilon}\land T})].

Setting xT,ε,𝐊=JT,ε,𝐊xX{\mathbb{Q}}^{T,\varepsilon,{\mathbf{K}}}_{x}=J_{T,\varepsilon,{\mathbf{K}}}\cdot{\mathbb{P}}_{x}^{X} completes the proof. ∎

8. Explosion and continuity at explosion

In this section we consider a QKS(θ,N)QKS(\theta,N)-process 𝕏{\mathbb{X}} with life-time ζ\zeta. We show that ζ=\zeta=\infty when θ(0,2)\theta\in(0,2) and that ζ<\zeta<\infty when θ2\theta\geq 2. In the latter case, we also prove that limtζXt\lim_{t\to\zeta-}X_{t} a.s. exists, for the usual topology of (2)N({\mathbb{R}}^{2})^{N}: the Keller-Segel process is continuous at explosion. This is not clear at all at first sight: we know that limtζXt=\lim_{t\to\zeta-}X_{t}=\triangle a.s. for the one-point compactification topology, which means that the process escapes from every compact of 𝒳{\mathcal{X}}, but it could either go to infinity, which is not difficult to exclude, or it could tend to the boundary of 𝒳{\mathcal{X}} without converging, e.g. because it could alternate very fast between having its particles labeled in [[1,k0]][\![1,k_{0}]\!] very close and having its particles labeled in [[2,k0+1]][\![2,k_{0}+1]\!] very close. The goal of the section is to prove the following result.

Proposition 8.1.

Fix θ>0\theta>0 and N2N\geq 2 such that N>θN>\theta, set k0=2N/θk_{0}=\lceil 2N/\theta\rceil and 𝒳=Ek0{\mathcal{X}}=E_{k_{0}} and consider a QKS(θ,N)QKS(\theta,N)-process 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳{}){\mathbb{X}}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}_{x}^{X})_{x\in{\mathcal{X}}\cup\{\triangle\}}) with life-time ζ\zeta. (i) If θ<2\theta<2, then quasi-everywhere, xX(ζ=)=1{\mathbb{P}}^{X}_{x}(\zeta=\infty)=1. (ii) If θ2\theta\geq 2, then quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., ζ<\zeta<\infty and Xζ=limtζXtX_{\zeta-}=\lim_{t\to\zeta}X_{t} exists for the usual topology of (2)N({\mathbb{R}}^{2})^{N} and does not belong to Ek0E_{k_{0}}.

We first show that the process does not explode in the subcritical case and cannot go to infinity at explosion in the supercritical case.

Lemma 8.2.

(i) If θ<2\theta<2 and N2N\geq 2, then quasi-everywhere, xX(ζ=)=1{\mathbb{P}}^{X}_{x}(\zeta=\infty)=1. (ii) If θ2\theta\geq 2 and N>θN>\theta, then quasi-everywhere,

xX(ζ< and supt[0,ζ)Xt<)=1.{\mathbb{P}}^{X}_{x}\Big{(}\zeta<\infty\hbox{ and }\color[rgb]{0,0,0}\sup_{t\in[0,\zeta)}\color[rgb]{0,0,0}\|X_{t}\|<\infty\Big{)}=1.
Proof.

The arguments below only apply quasi-everywhere, since we use Proposition 5.1. In both cases, we have for all i[[1,N]]i\in[\![1,N]\!] and all t[0,ζ)t\in[0,\zeta),

Xt22i=1N(XtiS[[1,N]](Xt)2+S[[1,N]](Xt)2)=2R[[1,N]](Xt)+2NS[[1,N]](Xt)2.||X_{t}||^{2}\leq 2\sum_{i=1}^{N}(\|X^{i}_{t}-S_{[\![1,N]\!]}(X_{t})\|^{2}+\|S_{[\![1,N]\!]}(X_{t})\|^{2})=2R_{[\![1,N]\!]}(X_{t})+2N\|S_{[\![1,N]\!]}(X_{t})\|^{2}.

By Lemma 5.2, there are a Brownian motion (Mt)t0(M_{t})_{t\geq 0} and a squared Bessel process (Dt)t0(D_{t})_{t\geq 0} with dimension dθ,N(N)d_{\theta,N}(N) (killed when it gets out of (0,)(0,\infty) if dθ,N(N)0d_{\theta,N}(N)\leq 0), such that S[[1,N]](Xt)=MtS_{[\![1,N]\!]}(X_{t})=M_{t} and R[[1,N]](Xt)=DtR_{[\![1,N]\!]}(X_{t})=D_{t} for all t[0,ζ)t\in[0,\zeta). These processes being locally bounded, we conclude that

(1) a.s., for all T>0,supt[0,ζT)Xt<.\displaystyle\mbox{ a.s., for all }T>0,\quad\sup_{t\in[0,\zeta\land T)}\|X_{t}\|<\infty.

(i) When θ<2\theta<2 and N2N\geq 2, we have k0=2N/θ>Nk_{0}=\lceil 2N/\theta\rceil>N, so that 𝒳=(2)N{\mathcal{X}}=({\mathbb{R}}^{2})^{N}. Hence on the event {ζ<}\{\zeta<\infty\}, we necessarily have lim suptζXt=\limsup_{t\to\zeta-}||X_{t}||=\infty, and this is incompatible with (1) with T=ζT=\zeta.

(ii) When N>θ2N>\theta\geq 2, we have dθ,N(N)0d_{\theta,N}(N)\leq 0, so that (Dt)t0(D_{t})_{t\geq 0} is killed at some finite time τ\tau. It holds that ζτ\zeta\leq\tau. Indeed, on the event where τ<ζ\tau<\zeta, we have R[[1,N]](Xτ)=limtτR[[1,N]](Xt)=limtτDt=0R_{[\![1,N]\!]}(X_{\tau})=\lim_{t\to\tau-}R_{[\![1,N]\!]}(X_{t})=\lim_{t\to\tau-}D_{t}=0, so that XτEk0X_{\tau}\notin E_{k_{0}} (since k0Nk_{0}\leq N), which is not possible since τ<ζ\tau<\zeta. Hence ζ\zeta is also a.s. finite and it holds that supt[0,ζ)Xt<\color[rgb]{0,0,0}\sup_{t\in[0,\zeta)}\color[rgb]{0,0,0}\|X_{t}\|<\infty a.s. by (1) with the choice T=ζT=\zeta. ∎

To show the continuity at explosion in the supercritical case, we need to prove the following delicate lemma.

Lemma 8.3.

Assume that N>θ2N>\theta\geq 2. Quasi-everywhere, for all K[[1,N]]K\subset[\![1,N]\!] with |K|2|K|\geq 2,

xX-a.s.,limtζRK(Xt)=0 or lim inftζRK(Xt)>0.{\mathbb{P}}^{X}_{x}\hbox{-\color[rgb]{0,0,0}a.s.\color[rgb]{0,0,0}},\qquad\lim_{t\to\zeta-}R_{K}(X_{t})=0\quad\mbox{ or }\quad\liminf_{t\to\zeta-}R_{K}(X_{t})>0.
Proof.

We proceed by reverse induction on the cardinal of KK. If first K=[[1,N]]K=[\![1,N]\!], the result is clear because (R[[1,N]](Xt))t[0,ζ)(R_{[\![1,N]\!]}(X_{t}))_{t\in[0,\zeta)} is a (killed) squared Bessel process on [0,ζ)[0,\zeta) by Lemma 5.2 (and since ζτ\zeta\leq\tau exactly as in the proof of Lemma 8.2-(ii)), hence it has a limit in +{\mathbb{R}}_{+} as tζt\to\zeta. Then, we assume that the property is proved if |K|n|K|\geq n where n[[3,N]]n\in[\![3,N]\!], we take K[[1,N]]K\subset[\![1,N]\!] such that |K|=n1|K|=n-1 and we show in several steps that a.s., either limtζRK(Xt)=0\lim_{t\to\zeta-}R_{K}(X_{t})=0 or lim inftζRK(Xt)>0\liminf_{t\to\zeta-}R_{K}(X_{t})>0.

Step 1. We fix ε(0,1]\varepsilon\in(0,1] and introduce σ~0ε=0\tilde{\sigma}^{\varepsilon}_{0}=0 and, for k1k\geq 1,

σkε=inf{t(σ~k1ε,ζ):RK(Xt)ε}andσ~kε=inf{t(σkε,ζ):RK(Xt)2ε},\sigma^{\varepsilon}_{k}=\inf\{t\in(\tilde{\sigma}^{\varepsilon}_{k-1},\zeta):R_{K}(X_{t})\leq\varepsilon\}\quad\hbox{and}\quad\tilde{\sigma}^{\varepsilon}_{k}=\inf\{t\in(\sigma^{\varepsilon}_{k},\zeta):R_{K}(X_{t})\geq 2\varepsilon\},

with the convention that inf=ζ\inf\emptyset=\zeta. We show in this step that for all deterministic A>0A>0, there exists a constant pA,ε>0p_{A,\varepsilon}>0 such that for all k1k\geq 1, quasi-everywhere, on {σkε<ζ}\{\sigma_{k}^{\varepsilon}<\zeta\},

xX({σ~kε(σkε+A)ζ}Bk,ε|σkεX)pA,ε,\displaystyle{\mathbb{P}}^{X}_{x}\Big{(}\{\tilde{\sigma}^{\varepsilon}_{k}\geq(\sigma^{\varepsilon}_{k}+A)\land\zeta\}\cup B_{k,\varepsilon}\Big{|}{\mathcal{M}}^{X}_{\sigma^{\varepsilon}_{k}}\Big{)}\geq p_{A,\varepsilon},

where tX=σ(Xs:s[0,t]){\mathcal{M}}^{X}_{t}=\sigma(X_{s}:s\in[0,t]), and where, setting aε=c|K|+1ε/c|K|a_{\varepsilon}=c_{|K|+1}\varepsilon/c_{|K|} (recall Lemma 6.2),

Bk,ε={supt[σkε,σ~kε)Xt1/ε or inft[σkε,σ~kε)miniKRK{i}(Xt)aε}.B_{k,\varepsilon}=\Big{\{}\color[rgb]{0,0,0}\sup_{t\in[\sigma^{\varepsilon}_{k},\tilde{\sigma}^{\varepsilon}_{k})}||X_{t}||\geq 1/\varepsilon\;\hbox{ or }\;\inf_{t\in[\sigma^{\varepsilon}_{k},\tilde{\sigma}^{\varepsilon}_{k})}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon}\Big{\}}.

By the strong Markov property of 𝕏{\mathbb{X}}, on {σkε<ζ}\{\sigma_{k}^{\varepsilon}<\zeta\},

xX({σ~kε(σkε+A)ζ}Bk,ε|σkεX)=g(Xσkε),{\mathbb{P}}^{X}_{x}\Big{(}\{\tilde{\sigma}^{\varepsilon}_{k}\geq(\sigma^{\varepsilon}_{k}+A)\land\zeta\}\cup B_{k,\varepsilon}\Big{|}{\mathcal{M}}^{X}_{\sigma^{\varepsilon}_{k}}\Big{)}=g(X_{\sigma_{k}^{\varepsilon}}),

where

g(y)=yX({σ~1ε(σ1ε+A)ζ}B1,ε)=yX({σ~1εAζ}C1,ε)g(y)={\mathbb{P}}^{X}_{y}\Big{(}\{\tilde{\sigma}^{\varepsilon}_{1}\geq(\sigma^{\varepsilon}_{1}+A)\land\zeta\}\cup B_{1,\varepsilon}\Big{)}={\mathbb{P}}^{X}_{y}\Big{(}\{\tilde{\sigma}^{\varepsilon}_{1}\geq A\land\zeta\}\cup C_{1,\varepsilon}\Big{)}

and

C1,ε={supt[0,σ~1ε)Xt1/ε or inft[0,σ~1ε)miniKRK{i}(Xt)aε}.C_{1,\varepsilon}=\Big{\{}\color[rgb]{0,0,0}\sup_{t\in[0,\tilde{\sigma}^{\varepsilon}_{1})}||X_{t}||\geq 1/\varepsilon\;\;\hbox{ or }\;\inf_{t\in[0,\tilde{\sigma}^{\varepsilon}_{1})}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon}\Big{\}}.

We used that RK(Xσkε)εR_{K}(X_{\sigma_{k}^{\varepsilon}})\leq\varepsilon on {σkε<ζ}\{\sigma_{k}^{\varepsilon}<\zeta\} by definition of σkε\sigma_{k}^{\varepsilon}, so that σ1ε=0\sigma^{\varepsilon}_{1}=0 under XσkεX{\mathbb{P}}^{X}_{X_{\sigma_{k}^{\varepsilon}}}. Using again that RK(Xσkε)εR_{K}(X_{\sigma_{k}^{\varepsilon}})\leq\varepsilon on {σkε<ζ}\{\sigma_{k}^{\varepsilon}<\zeta\}, it suffices to show that there is a constant pA,ε>0p_{A,\varepsilon}>0 such that g(y)pA,εg(y)\geq p_{A,\varepsilon} quasi-everywhere in {y𝒳:RK(y)ε}\{y\in{\mathcal{X}}:R_{K}(y)\leq\varepsilon\}.

If first y1/ε||y||\geq 1/\varepsilon or miniKRK{i}(y)aε\min_{i\notin K}R_{K\cup\{i\}}(y)\leq a_{\varepsilon}, then clearly, g(y)=1g(y)=1.

Otherwise, yG𝐊,εy\in G_{{\mathbf{K}},\varepsilon}, where

G𝐊,ε={x𝒳: for all iK, all jK,xixj2>ε}B(0,1/ε)G_{{\mathbf{K}},\varepsilon}=\{x\in{\mathcal{X}}:\mbox{ for all }i\in K,\mbox{ all }j\notin K,\;\|x^{i}-x^{j}\|^{2}>\varepsilon\}\cap B(0,1/\varepsilon)

as in Proposition 7.1 with 𝐊=(K,Kc){\mathbf{K}}=(K,K^{c}), because y<1/ε||y||<1/\varepsilon and because RK(y)ε<2εR_{K}(y)\leq\varepsilon<2\varepsilon and miniKRK{i}(y)>aε=c|K|+1ε/c|K|\min_{i\notin K}R_{K\cup\{i\}}(y)>a_{\varepsilon}=c_{|K|+1}\varepsilon/c_{|K|} imply that xixk2>ε||x^{i}-x^{k}||^{2}>\varepsilon for all iKi\in K, jKj\notin K by Lemma 6.2. For the very same reasons and by definition of σ~1ε\tilde{\sigma}_{1}^{\varepsilon}, it holds that

(2) C1,εc{for all t[0,σ~1ε),XtG𝐊,ε}.C_{1,\varepsilon}^{c}\subset\{\hbox{for all }t\in[0,\tilde{\sigma}_{1}^{\varepsilon}),\;\;X_{t}\in G_{{\mathbf{K}},\varepsilon}\}.

We now apply Proposition 7.1 with T=AT=A (and ε\varepsilon) and we find that quasi-everywhere in G𝐊,εG_{{\mathbf{K}},\varepsilon},

g(y)\displaystyle g(y)\geq CA,ε,𝐊1yA,ε,𝐊({σ~1εAζ}C1,ε)\displaystyle C_{A,\varepsilon,{\mathbf{K}}}^{-1}{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}(\{\tilde{\sigma}^{\varepsilon}_{1}\geq A\land\zeta\}\cup C_{1,\varepsilon})
(3) =\displaystyle= CA,ε,𝐊1yA,ε,𝐊({σ~1εAζ}C1,εc)+CA,ε,𝐊1yA,ε,𝐊(C1,ε).\displaystyle C_{A,\varepsilon,{\mathbf{K}}}^{-1}{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}(\{\tilde{\sigma}^{\varepsilon}_{1}\geq A\land\zeta\}\cap C_{1,\varepsilon}^{c})+C_{A,\varepsilon,{\mathbf{K}}}^{-1}{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}(C_{1,\varepsilon}).

But we know from Proposition 7.1 and Lemma 5.2 that under yA,ε,𝐊{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}, (RK(Xt))t[0,τK,εA](R_{K}(X_{t}))_{t\in[0,\tau_{K,\varepsilon}\land A]} is a squared Bessel process with dimension dθ|K|/N,|K|(|K|)=dθ,N(|K|)d_{\theta|K|/N,|K|}(|K|)=d_{\theta,N}(|K|), issued from RK(y)εR_{K}(y)\leq\varepsilon, stopped at time τ𝐊,εA\tau_{{\mathbf{K}},\varepsilon}\land A, where τ𝐊,ε=inf{t>0:XtG𝐊,ε}\tau_{{\mathbf{K}},\varepsilon}=\inf\{t>0:X_{t}\notin G_{{\mathbf{K}},\varepsilon}\}. Hence there exists, under yA,ε,𝐊{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}, a squared Bessel process (St)t0(S_{t})_{t\geq 0} with dimension dθ,N(|K|)d_{\theta,N}(|K|) such that St=RK(Xt)S_{t}=R_{K}(X_{t}) for all t[0,τ𝐊,εA]t\in[0,\tau_{{\mathbf{K}},\varepsilon}\land A]. We introduce κε=inf{t>0:St2ε}\kappa_{\varepsilon}=\inf\{t>0:S_{t}\geq 2\varepsilon\} and we observe that

{κεAζ}C1,εc={σ~1εA}C1,εc.\displaystyle\{\kappa_{\varepsilon}\geq A\land\zeta\}\cap C_{1,\varepsilon}^{c}=\{\tilde{\sigma}^{\varepsilon}_{1}\geq A\}\cap C_{1,\varepsilon}^{c}.

Indeed, we used that on C1,εcC_{1,\varepsilon}^{c}, we have τ𝐊,εσ~1ε\tau_{{\mathbf{K}},\varepsilon}\geq\tilde{\sigma}^{\varepsilon}_{1} by (2) so that RK(Xt)=StR_{K}(X_{t})=S_{t} for all t[0,σ~1εA)t\in[0,\tilde{\sigma}^{\varepsilon}_{1}\land A), from which we conclude that κεAζ\kappa_{\varepsilon}\geq A\land\zeta if and only σ~1εAζ\tilde{\sigma}^{\varepsilon}_{1}\geq A\land\zeta. Coming back to (3), we get

g(y)\displaystyle g(y)\geq CA,ε,𝐊1yA,ε,𝐊({κεAζ}C1,εc)+CA,ε,𝐊1yA,ε,𝐊(C1,ε)=CA,ε,𝐊1yA,ε,𝐊(κεAζ).\displaystyle C_{A,\varepsilon,{\mathbf{K}}}^{-1}{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}(\{\kappa_{\varepsilon}\geq A\land\zeta\}\cap C_{1,\varepsilon}^{c})+C_{A,\varepsilon,{\mathbf{K}}}^{-1}{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}(C_{1,\varepsilon})=C_{A,\varepsilon,{\mathbf{K}}}^{-1}{\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}(\kappa_{\varepsilon}\geq A\land\zeta).

The step is complete, since yA,ε,𝐊(κεA){\mathbb{Q}}^{A,\varepsilon,{\mathbf{K}}}_{y}(\kappa_{\varepsilon}\geq A) is the probability that a squared Bessel process with dimension dθ,N(|K|)d_{\theta,N}(|K|) issued from RK(y)εR_{K}(y)\leq\varepsilon remains below 2ε2\varepsilon during [0,A][0,A] and is thus strictly positive, uniformly in yy (such that yG𝐊,εy\in G_{{\mathbf{K}},\varepsilon} and RK(y)εR_{K}(y)\leq\varepsilon).

Step 2. We prove here that for all ε(0,1]\varepsilon\in(0,1], all A>0A>0, quasi-everywhere,

xX(lim suptζ||Xt||1/ε or lim inftζminiKRK{i}(Xt)aε or k1,σkεζA)=1.{\mathbb{P}}^{X}_{x}\Big{(}\limsup_{t\to\zeta-}||X_{t}||\geq 1/\varepsilon\;\;\hbox{ or }\;\;\liminf_{t\to\zeta-}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon}\;\;\hbox{ or }\;\;\exists\;k\geq 1,\;\sigma^{\varepsilon}_{k}\geq\zeta\land A\Big{)}=1.

All the arguments below only hold quasi-everywhere, even if we do not mention it explicitly during this step. For k1k\geq 1, we introduce, with Bk,εB_{k,\varepsilon} defined in Step 1,

Ωk+1={σk+1ε<ζA}Bk,εc\Omega_{k+1}=\{\sigma^{\varepsilon}_{k+1}<\zeta\land A\}\cap B_{k,\varepsilon}^{c}

and we first show that xX(lim infkΩk)=0{\mathbb{P}}^{X}_{x}(\liminf_{k}\Omega_{k})=0. To this end, it suffices to check that for all 1\ell\geq 1, xX(k=Ωk)=0{\mathbb{P}}^{X}_{x}(\cap_{k=\ell}^{\infty}\Omega_{k})=0. Since Ωk\Omega_{k} is σkε{\mathcal{M}}_{\sigma^{\varepsilon}_{k}}-measurable, for all m1m\geq\ell\geq 1,

xX(k=m+1Ωk)=𝔼xX[1Ik=mΩkxX(Ωm+1|σmε)].{\mathbb{P}}^{X}_{x}(\cap_{k=\ell}^{m+1}\Omega_{k})={\mathbb{E}}^{X}_{x}[\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\cap_{k=\ell}^{m}\Omega_{k}}{\mathbb{P}}^{X}_{x}(\Omega_{m+1}|{\mathcal{M}}_{\sigma^{\varepsilon}_{m}})].

Since moreover k=mΩk{σmε<ζ}\cap_{k=\ell}^{m}\Omega_{k}\subset\{\sigma^{\varepsilon}_{m}<\zeta\} and since σm+1εσ~mεσ~mεσmε\sigma^{\varepsilon}_{m+1}\geq\tilde{\sigma}^{\varepsilon}_{m}\geq\tilde{\sigma}^{\varepsilon}_{m}-\sigma_{m}^{\varepsilon}, we deduce that on k=mΩk\cap_{k=\ell}^{m}\Omega_{k},

xX(Ωm+1|σmε)=\displaystyle{\mathbb{P}}^{X}_{x}(\Omega_{m+1}|{\mathcal{M}}_{\sigma^{\varepsilon}_{m}})= 1xX({σm+1εζA}Bm,ε|σmε)\displaystyle 1-{\mathbb{P}}^{X}_{x}(\{\sigma^{\varepsilon}_{m+1}\geq\zeta\land A\}\cup B_{m,\varepsilon}|{\mathcal{M}}_{\sigma^{\varepsilon}_{m}})
\displaystyle\leq 1xX({σ~mε(σmε+A)ζ}Bm,ε|σmε),\displaystyle 1-{\mathbb{P}}^{X}_{x}(\{\tilde{\sigma}^{\varepsilon}_{m}\geq(\sigma^{\varepsilon}_{m}+A)\land\zeta\}\cup B_{m,\varepsilon}|{\mathcal{M}}_{\sigma^{\varepsilon}_{m}}),

so that xX(Ωm+1|σmε)1pA,ε{\mathbb{P}}^{X}_{x}(\Omega_{m+1}|{\mathcal{M}}_{\sigma^{\varepsilon}_{m}})\leq 1-p_{A,\varepsilon} by Step 1. Hence we conclude that

xX(k=m+1Ωk)(1pA,ε)xX(k=mΩk){\mathbb{P}}^{X}_{x}(\cap_{k=\ell}^{m+1}\Omega_{k})\leq(1-p_{A,\varepsilon}){\mathbb{P}}^{X}_{x}(\cap_{k=\ell}^{m}\Omega_{k})

for all m1m\geq\ell\geq 1, so that xX(k=Ωk)=0{\mathbb{P}}^{X}_{x}(\cap_{k=\ell}^{\infty}\Omega_{k})=0 as desired.

Hence xX(lim infkΩk)=0{\mathbb{P}}^{X}_{x}(\liminf_{k}\Omega_{k})=0, so that a.s., an infinite number of Ωkc\Omega_{k}^{c} are realized. Recalling that

Ωk+1c={σk+1εζA or inft[σkε,σ~kε)miniKRK{i}(Xt)aε or supt[σkε,σ~kε)Xt1/ε},\Omega_{k+1}^{c}=\Big{\{}\sigma^{\varepsilon}_{k+1}\geq\zeta\land A\;\;\hbox{ or }\;\;\inf_{t\in[\sigma^{\varepsilon}_{k},\tilde{\sigma}^{\varepsilon}_{k})}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon}\;\;\hbox{ or }\;\;\sup_{t\in[\sigma^{\varepsilon}_{k},\tilde{\sigma}^{\varepsilon}_{k})}||X_{t}||\geq 1/\varepsilon\Big{\}},

we find the following alternative:

\bullet either there is k1k\geq 1 such that σkεζA\sigma^{\varepsilon}_{k}\geq\zeta\land A; \bullet or for all k1k\geq 1, σkε<ζ\sigma^{\varepsilon}_{k}<\zeta and inft[σkε,σ~kε)miniKRK{i}(Xt)aε\color[rgb]{0,0,0}\inf_{t\in[\sigma^{\varepsilon}_{k},\tilde{\sigma}^{\varepsilon}_{k})}\color[rgb]{0,0,0}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon} for infinitely many kk’s, which implies that lim inftζminiKRK{i}(Xt)aε\liminf_{t\to\zeta-}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon} because necessarily, limσkε=ζ\lim_{\infty}\sigma^{\varepsilon}_{k}=\zeta by definition of the sequence (σkε)k1(\sigma_{k}^{\varepsilon})_{k\geq 1} and by continuity of tRK(Xt)t\to R_{K}(X_{t}) on [0,ζ)[0,\zeta); \bullet or for all k1k\geq 1, σkε<ζ\sigma^{\varepsilon}_{k}<\zeta and there are infinitely many kk’s for which supt[σkε,σ~kε)Xt1/ε\sup_{t\in[\sigma^{\varepsilon}_{k},\tilde{\sigma}^{\varepsilon}_{k})}||X_{t}||\geq 1/\varepsilon and this implies that lim suptζXt1/ε\limsup_{t\to\zeta-}||X_{t}||\geq 1/\varepsilon, because limσkε=ζ\lim_{\infty}\sigma^{\varepsilon}_{k}=\zeta as previously.

Step 3. We conclude. Applying Step 2, we find that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., for all AA\in{\mathbb{N}} and all ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1],

lim suptζXt1/ε or lim inftζminiKRK{i}(Xt)aε or k1,σkεζA.\limsup_{t\to\zeta-}||X_{t}||\geq 1/\varepsilon\;\;\hbox{ or }\;\;\liminf_{t\to\zeta-}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon}\;\;\hbox{ or }\;\;\exists\;k\geq 1,\;\sigma^{\varepsilon}_{k}\geq\zeta\land A.

By Lemma 8.2-(ii), we know that ζ<\zeta<\infty, so that choosing A=ζA=\lceil\zeta\rceil, we conclude that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., for all ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1]

(4) lim suptζXt1/ε or lim inftζminiKRK{i}(Xt)aε or k1,σkε=ζ.\limsup_{t\to\zeta-}||X_{t}||\geq 1/\varepsilon\;\;\hbox{ or }\;\;\liminf_{t\to\zeta-}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})\leq a_{\varepsilon}\;\;\hbox{ or }\;\;\exists\;k\geq 1,\;\sigma^{\varepsilon}_{k}=\zeta.

And by Lemma 8.2-(ii) again, lim suptζXt1/ε0\limsup_{t\to\zeta-}||X_{t}||\leq 1/\varepsilon_{0} for some (random) ε0(0,1]\varepsilon_{0}\in(0,1].

On the event where lim inftζminiKRK{i}(Xt)=0\liminf_{t\to\zeta-}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})=0, there exists some (random) i0Ki_{0}\notin K such that lim inftζRK{i0}(Xt)=0\liminf_{t\to\zeta-}R_{K\cup\{i_{0}\}}(X_{t})=0, whence limtζRK{i0}(Xt)=0\lim_{t\to\zeta-}R_{K\cup\{i_{0}\}}(X_{t})=0 by induction assumption, and this obviously implies that limtζRK(Xt)=0\lim_{t\to\zeta-}R_{K}(X_{t})=0.

On the complementary event, we fix ε1(0,ε0]\varepsilon_{1}\in(0,\varepsilon_{0}] such that lim inftζminiKRK{i}(Xt)>aε1\liminf_{t\to\zeta-}\min_{i\notin K}R_{K\cup\{i\}}(X_{t})>a_{\varepsilon_{1}} and we conclude from (4) and the fact that lim suptζXt1/ε0\limsup_{t\to\zeta-}||X_{t}||\leq 1/\varepsilon_{0} that for all ε(0,ε1]\varepsilon\in{\mathbb{Q}}\cap(0,\varepsilon_{1}], there exists kε1k_{\varepsilon}\geq 1 such that σkεε=ζ\sigma_{k_{\varepsilon}}^{\varepsilon}=\zeta. Recalling the definition of (σkε)k1(\sigma^{\varepsilon}_{k})_{k\geq 1}, we deduce that for all ε(0,ε1]\varepsilon\in{\mathbb{Q}}\cap(0,\varepsilon_{1}], RK(Xt)R_{K}(X_{t}) upcrosses the segment [ε,2ε][\varepsilon,2\varepsilon] a finite number of times during [0,ζ)[0,\zeta). Hence for all ε(0,ε1]\varepsilon\in(0,\varepsilon_{1}]\cap{\mathbb{Q}}, there exists tε[0,ζ)t_{\varepsilon}\in[0,\zeta) such that either RK(Xt)>εR_{K}(X_{t})>\varepsilon for all t[tε,ζ)t\in[t_{\varepsilon},\zeta) or RK(Xt)<2εR_{K}(X_{t})<2\varepsilon for all t[tε,ζ)t\in[t_{\varepsilon},\zeta). If there is ε(0,ε1]\varepsilon\in{\mathbb{Q}}\cap(0,\varepsilon_{1}] such that RK(Xt)>εR_{K}(X_{t})>\varepsilon for all t[tε,ζ)t\in[t_{\varepsilon},\zeta), then lim inftζRK(Xt)ε>0\liminf_{t\to\zeta-}R_{K}(X_{t})\geq\varepsilon>0. If next for all ε(0,ε1]\varepsilon\in{\mathbb{Q}}\cap(0,\varepsilon_{1}], we have RK(Xt)<2εR_{K}(X_{t})<2\varepsilon for all t[tε,ζ)t\in[t_{\varepsilon},\zeta), then limtζRK(Xt)=0\lim_{t\to\zeta-}R_{K}(X_{t})=0.

Hence in any case, we have either limtζRK(Xt)=0\lim_{t\to\zeta-}R_{K}(X_{t})=0 or lim inftζRK(Xt)>0\liminf_{t\to\zeta-}R_{K}(X_{t})>0. ∎

We finally give the

Proof of Proposition 8.1.

Point (i), which concerns the subcritical case, has already been checked in Lemma 8.2-(i). Concerning point (ii), which concerns the supercritical case θ2\theta\geq 2, we already know that quasi-everywhere, xX(ζ<)=1{\mathbb{P}}^{X}_{x}(\zeta<\infty)=1 by Lemma 8.2-(ii), and it remains to prove that xX{\mathbb{P}}^{X}_{x}-a.s., limtζXt\lim_{t\rightarrow\zeta-}X_{t} exists and does not belong to Ek0E_{k_{0}}. We divide the proof in four steps.

Step 1. For 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} a partition of [[1,N]][\![1,N]\!] and ε(0,1]\varepsilon\in(0,1], we consider as in Proposition 7.1

G𝐊,ε={x𝒳:min1pqminiKp,jKqxixj2>ε}B(0,1ε)G_{{\mathbf{K}},\varepsilon}=\Big{\{}x\in{\mathcal{X}}:\min_{1\leq p\neq q\leq\ell}\;\;\min_{i\in K_{p},j\in K_{q}}\|x^{i}-x^{j}\|^{2}>\varepsilon\Big{\}}\cap B\Big{(}0,\frac{1}{\varepsilon}\Big{)}

and τ𝐊,ε=inf{t0:XtG𝐊,ε}[0,ζ]\tau_{{\mathbf{K}},\varepsilon}=\inf\{t\geq 0:X_{t}\notin G_{{\mathbf{K}},\varepsilon}\}\in[0,\zeta]. We show here for each T>0T>0, quasi-everywhere in G𝐊,εG_{{\mathbf{K}},\varepsilon}, xX{\mathbb{P}}^{X}_{x}-a.s., for all T>0T>0, all p[[1,]]p\in[\![1,\ell]\!], SKp(Xt)S_{K_{p}}(X_{t}) has a limit in 2{\mathbb{R}}^{2} as t(τ𝐊,εT)t\to(\tau_{{\mathbf{K}},\varepsilon}\land T)-.

If =1\ell=1, the result is obvious since S[[1,N]](Xt)S_{[\![1,N]\!]}(X_{t}) is a Brownian motion during [0,ζ)[0,\zeta) by Lemma 5.2. If next 2\ell\geq 2, Proposition 7.1 and Lemma 5.2 tell us that under xT,ε,𝐊{\mathbb{Q}}_{x}^{T,\varepsilon,{\mathbf{K}}}, which is equivalent to xX{\mathbb{P}}^{X}_{x}, the processes SKp(Xt)S_{K_{p}}(X_{t}) are some Brownian motions on [0,τ𝐊,εT)[0,\tau_{{\mathbf{K}},\varepsilon}\land T), and thus have some limits as t(τ𝐊,εT)t\to(\tau_{{\mathbf{K}},\varepsilon}\land T)-.

Step 2. For ε(0,1]\varepsilon\in(0,1] and 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} a partition of [[1,N]][\![1,N]\!], we set η~0𝐊,ε=0\tilde{\eta}_{0}^{{\mathbf{K}},\varepsilon}=0 and, for k0k\geq 0,

ηk+1𝐊,ε=inf{tη~k𝐊,ε:XtG𝐊,2ε} and η~k+1𝐊,ε=inf{tηk+1𝐊,ε:XtG𝐊,ε},\eta_{k+1}^{{\mathbf{K}},\varepsilon}=\inf\{t\geq\tilde{\eta}_{k}^{{\mathbf{K}},\varepsilon}:X_{t}\in G_{{\mathbf{K}},2\varepsilon}\}\quad\mbox{ and }\quad\tilde{\eta}_{k+1}^{{\mathbf{K}},\varepsilon}=\inf\{t\geq\eta_{k+1}^{{\mathbf{K}},\varepsilon}:X_{t}\notin G_{{\mathbf{K}},\varepsilon}\},

with the convention that inf=ζ\inf\emptyset=\zeta. Using Step 1 and the strong Markov property, we conclude that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., for all ε(0,1]\varepsilon\in(0,1]\cap{\mathbb{Q}}, all k1k\geq 1, all T+T\in{\mathbb{N}}_{+}, on {ηk𝐊,ε<ζ}\{\eta_{k}^{{\mathbf{K}},\varepsilon}<\zeta\}, for all p[[1,]]p\in[\![1,\ell]\!], SKp(Xt)S_{K_{p}}(X_{t}) admits a limit in 2{\mathbb{R}}^{2} as tt goes to (η~k𝐊,εT)(\tilde{\eta}^{{\mathbf{K}},\varepsilon}_{k}\land T)-. Choosing T=ζT=\lceil\zeta\rceil, we conclude that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., on {ηk𝐊,ε<ζ}\{\eta_{k}^{{\mathbf{K}},\varepsilon}<\zeta\}, for all ε(0,1]\varepsilon\in(0,1]\cap{\mathbb{Q}}, all k1k\geq 1, all p[[1,]]p\in[\![1,\ell]\!],

SKp(Xt) admits a limit in 2 as t goes to η~k𝐊,ε.S_{K_{p}}(X_{t})\mbox{ admits a limit in ${\mathbb{R}}^{2}$ as }t\mbox{ goes to }\tilde{\eta}^{{\mathbf{K}},\varepsilon}_{k}-.

Step 3. We now check that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., there is a partition 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} of [[1,N]][\![1,N]\!], some ε(0,1]\varepsilon\in(0,1]\cap{\mathbb{Q}} and some k1k\geq 1 such that (i) ηk𝐊,ε<ζ\eta_{k}^{{\mathbf{K}},\varepsilon}<\zeta and η~k𝐊,ε=ζ\tilde{\eta}^{{\mathbf{K}},\varepsilon}_{k}=\zeta and (ii) for all p[[1,]]p\in[\![1,\ell]\!], limtζRKp(Xt)=0\lim_{t\to\zeta-}R_{K_{p}}(X_{t})=0.

By Lemma 8.3, we know that for all K[[1,N]]K\subset[\![1,N]\!], we have the alternative limtζRK(Xt)=0\lim_{t\to\zeta-}R_{K}(X_{t})=0 or lim inftζRK(Xt)>0\liminf_{t\to\zeta-}R_{K}(X_{t})>0. Hence the partition 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} of [[1,N]][\![1,N]\!] consisting of the classes of the equivalence relation defined by iji\sim j if and only if limtζR{i,j}(Xt)=0\lim_{t\to\zeta}R_{\{i,j\}}(X_{t})=0 satisfies that for all p[[1,]]p\in[\![1,\ell]\!], limtζRKp(Xt)=0\lim_{t\to\zeta-}R_{K_{p}}(X_{t})=0 and lim inftζminiKpRKp{i}(Xt)>0\liminf_{t\to\zeta-}\min_{i\notin K_{p}}R_{K_{p}\cup\{i\}}(X_{t})>0.

Using moreover that lim suptζXt<\limsup_{t\to\zeta-}||X_{t}||<\infty according to Lemma 8.2, we deduce that there is α(0,ζ)\alpha\in(0,\zeta) and ε(0,1]\varepsilon\in(0,1]\cap{\mathbb{Q}} such that for all t[α,ζ)t\in[\alpha,\zeta), XtX_{t} belongs to G𝐊,2εG_{{\mathbf{K}},2\varepsilon}. Finally, we consider k=max{m1:ηm𝐊,εα}k=\max\{m\geq 1:\eta^{{\mathbf{K}},\varepsilon}_{m}\leq\alpha\}, which is finite by continuity of tXtt\mapsto X_{t} on [0,α][0,\alpha], and it holds that ηk𝐊,εα<ζ\eta^{{\mathbf{K}},\varepsilon}_{k}\leq\alpha<\zeta and that η~k𝐊,ε=ζ\tilde{\eta}^{{\mathbf{K}},\varepsilon}_{k}=\zeta.

Step 4. We consider the (random) partition 𝐊=(Kp)p[[1,]]{\mathbf{K}}=(K_{p})_{p\in[\![1,\ell]\!]} introduced in Step 3. By Step 2 and since ηk𝐊,ε<ζ\eta_{k}^{{\mathbf{K}},\varepsilon}<\zeta and η~k𝐊,ε=ζ\tilde{\eta}^{{\mathbf{K}},\varepsilon}_{k}=\zeta, we know that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., for all p[[1,]]p\in[\![1,\ell]\!], Mp=limtζSKp(Xt)M_{p}=\lim_{t\to\zeta-}S_{K_{p}}(X_{t}) exists in 2{\mathbb{R}}^{2}. By Step 3, we know that for all p[[1,]]p\in[\![1,\ell]\!], limtζRKp(Xt)=0\lim_{t\to\zeta}R_{K_{p}}(X_{t})=0. We easily conclude that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., for all p[[1,]]p\in[\![1,\ell]\!], all iKpi\in K_{p}, limtζXti=Mp\lim_{t\to\zeta-}X^{i}_{t}=M_{p}. This shows that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., Xζ=limtζXtX_{\zeta-}=\lim_{t\to\zeta-}X_{t} exists in (2)N({\mathbb{R}}^{2})^{N}. Moreover, XζX_{\zeta-} cannot belong to 𝒳=Ek0{\mathcal{X}}=E_{k_{0}}, because limtζXt=\lim_{t\to\zeta-}X_{t}=\triangle when Ek0{}E_{k_{0}}\cup\{\triangle\} is endowed with the one-point compactification topology, see Subsection B.1. ∎

9. Some special cases

During a KK-collision, the particles labeled in KK are isolated from the other ones. Thanks to Proposition 7.1, it will thus be possible to describe what happens in a neighborhood of the instant of this KK-collision, by studying a QKS(θ|K|/N,|K|)QKS(\theta|K|/N,|K|)-process. In other words, we may assume that |K|=N|K|=N, so that the following special cases, which are the purpose of this section, will be crucial.

Proposition 9.1.

Let N4N\geq 4 and θ>0\theta>0 such that N>θN>\theta. Consider a QKS(θ,N)QKS(\theta,N)-process 𝕏{\mathbb{X}} as in Proposition 4.1. Recall that ζ=inf{t0:Xt=}\zeta=\inf\{t\geq 0:X_{t}=\triangle\} and set τ=inf{t0:R[[1,N]](Xt)(0,)}\tau=\inf\{t\geq 0:R_{[\![1,N]\!]}(X_{t})\notin(0,\infty)\} with the convention that RK()=0R_{K}(\triangle)=0, so that τ[0,ζ]\tau\in[0,\zeta].

(i) If dθ,N(N1)0d_{\theta,N}(N-1)\leq 0 and dθ,N(N)<2d_{\theta,N}(N)<2, then quasi-everywhere,

xX(inft[0,ζ)R[[1,N]](Xt)>0)=1.{\mathbb{P}}^{X}_{x}\Big{(}\inf_{t\in[0,\zeta)}R_{[\![1,N]\!]}(X_{t})>0\Big{)}=1.

(ii) If dθ,N(N1)(0,2)d_{\theta,N}(N-1)\in(0,2) and dθ,N(N)<2d_{\theta,N}(N)<2, then quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s, for all K[[1,N]]K\subset[\![1,N]\!] with cardinal |K|=N1|K|=N-1, there is t[0,τ)t\in[0,\tau) such that RK(Xt)=0R_{K}(X_{t})=0.

(iii) If 0<dθ,N(N)<2dθ,N(N1)0<d_{\theta,N}(N)<2\leq d_{\theta,N}(N-1), then quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s, for all K[[1,N]]K\subset[\![1,N]\!] with cardinal |K|=2|K|=2, there is t[0,τ)t\in[0,\tau) such that RK(Xt)=0R_{K}(X_{t})=0.

The proof of this proposition is very long. First, we recall some notation about the decomposition of 𝕏{\mathbb{X}} obtained in Proposition 5.1 and we study the involved time-change. We then derive a formula describing RK(Ut)R_{K}(U_{t}), valid on certain time intervals, for any K[[1,N]]K\subset[\![1,N]\!]. This formula is of course not closed, but it allows us to compare RK(Ut)R_{K}(U_{t}), when it is close to 0, to some process resembling a squared Bessel process, of which one easily studies the behavior near 0. Finally, we prove Proposition 9.1, unifying a little points (i) and (ii) and treating separately point (iii).

9.1. Notation and preliminaries

We recall the decomposition of Proposition 5.1, which holds true quasi-everywhere in 𝒳EN{\mathcal{X}}\cap E_{N}. Consider a Brownian motion (Mt)t0(M_{t})_{t\geq 0} with diffusion coefficient N1/2N^{-1/2} starting from S[[1,N]](x)S_{[\![1,N]\!]}(x), a squared Bessel process (Dt)t0(D_{t})_{t\geq 0} starting from R[[1,N]](x)>0R_{[\![1,N]\!]}(x)>0 killed when leaving (0,)(0,\infty) with life-time τD=inf{t0:Dt=}\tau_{D}=\inf\{t\geq 0:D_{t}=\triangle\} and a QSKS(θ,N)QSKS(\theta,N) -process (Ut)t0(U_{t})_{t\geq 0} starting from Φ𝕊(x)\Phi_{\mathbb{S}}(x) with life-time ξ=inf{t0:Ut=}\xi=\inf\{t\geq 0:U_{t}=\triangle\}, all these processes being independent. For t[0,τD)t\in[0,\tau_{D}), we put At=0tdsDsA_{t}=\int_{0}^{t}\frac{{\rm d}s}{D_{s}}. We also consider the inverse ρ:[0,AτD)[0,τD)\rho:[0,A_{\tau_{D}})\to[0,\tau_{D}) of AA.

Lemma 9.2.

If dθ,N(N)<2d_{\theta,N}(N)<2, then τD<\tau_{D}<\infty and AτD=A_{\tau_{D}}=\infty a.s.

Proof.

Since (Dt)t0(D_{t})_{t\geq 0} is a (killed) squared Bessel process with dimension dθ,N(N)<2d_{\theta,N}(N)<2, we have τD<\tau_{D}<\infty a.s according to Revuz-Yor [21, Chapter XI]. Moreover, there is a Brownian motion (Bt)t0(B_{t})_{t\geq 0} such that Dt=r+20tDsdBs+dθ,N(N)tD_{t}=r+2\int_{0}^{t}\sqrt{D_{s}}{\rm d}B_{s}+d_{\theta,N}(N)t for all t[0,τD)t\in[0,\tau_{D}), where r=R[[1,N]](x)>0r=R_{[\![1,N]\!]}(x)>0. A simple computation shows the existence of a Brownian motion (Wt)t0(W_{t})_{t\geq 0} such that for all t[0,AτD)t\in[0,A_{\tau_{D}}),

Dρt=r+20tDρsdWs+dθ,N(N)0tDρsds.D_{\rho_{t}}=r+2\int_{0}^{t}D_{\rho_{s}}{\rm d}W_{s}+d_{\theta,N}(N)\int_{0}^{t}D_{\rho_{s}}{\rm d}s.

Hence for all t[0,AτD)t\in[0,A_{\tau_{D}}), Dρt=rexp(2Wt+(dθ,N(N)2)t)D_{\rho_{t}}=r\exp(2W_{t}+(d_{\theta,N}(N)-2)t). On the event where AτD<A_{\tau_{D}}<\infty, we have 0=DτD=limtAτDDρt=exp(2WAτD+(dθ,N(N)2)AτD)>00=D_{\tau_{D}-}=\lim_{t\to A_{\tau_{D}}}D_{\rho_{t}}=\exp(2W_{A_{\tau_{D}}}+(d_{\theta,N}(N)-2)A_{\tau_{D}})>0. Hence AτD=A_{\tau_{D}}=\infty a.s. ∎

From now on, we assume that dθ,N(N)<2d_{\theta,N}(N)<2. Hence A:[0,τD)[0,)A:[0,\tau_{D})\to[0,\infty) is an increasing bijection, as well as ρ:[0,)[0,τD)\rho:[0,\infty)\to[0,\tau_{D}). By Proposition 5.1, quasi-everywhere in 𝒳EN{\mathcal{X}}\cap E_{N} , we can find a triple (Mt,Dt,Ut)t0(M_{t},D_{t},U_{t})_{t\geq 0} as above such that for 𝕏{\mathbb{X}} our QKS(θ,N)QKS(\theta,N) process starting from xx, for all t[0,τDρξ)t\in[0,\tau_{D}\land\rho_{\xi}), and actually for all t[0,ρξ)t\in[0,\rho_{\xi}) because ρξτD\rho_{\xi}\leq\tau_{D} since ρ\rho is [0,τD)[0,\tau_{D})-valued,

Xt=Ψ(Mt,Dt,UAt),i.e.Mt=S[[1,N]](Xt),Dt=R[[1,N]](Xt)andUAt=Φ𝕊(Xt).X_{t}=\Psi(M_{t},D_{t},U_{A_{t}}),\quad\hbox{i.e.}\quad M_{t}=S_{[\![1,N]\!]}(X_{t}),\quad D_{t}=R_{[\![1,N]\!]}(X_{t})\quad\hbox{and}\quad U_{A_{t}}=\Phi_{\mathbb{S}}(X_{t}).

We recall that Ψ(m,r,u)=γ(m)+ru\Psi(m,r,u)=\gamma(m)+\sqrt{r}u if (m,r,u)2×(0,)×𝒰(m,r,u)\in{\mathbb{R}}^{2}\times(0,\infty)\times{\mathcal{U}} and Ψ(m,r,u)=\Psi(m,r,u)=\triangle if (m,r,u)=(m,r,u)=\triangle. Observe that τ=τDρξ=ρξ\tau=\tau_{D}\land\rho_{\xi}=\rho_{\xi}, where τ=inf{t0:R[[1,N]](Xt)(0,)}[0,ζ]\tau=\inf\{t\geq 0:R_{[\![1,N]\!]}(X_{t})\notin(0,\infty)\}\in[0,\zeta].

We note that if ξ<\xi<\infty, then ρξ<τD\rho_{\xi}<\tau_{D}, because ρ\rho is an increasing bijection from [0,)[0,\infty) into [0,τD)[0,\tau_{D}). Hence, still if ξ<\xi<\infty, then XX explodes at time ρξ\rho_{\xi} strictly before τD\tau_{D}, whence

(1) {ξ<}{inft[0,ζ)R[[1,N]](Xt)>0}.\{\xi<\infty\}\subset\Big{\{}\inf_{t\in[0,\zeta)}R_{[\![1,N]\!]}(X_{t})>0\Big{\}}.

Finally note that since UU is 𝕊{\mathbb{S}}-valued, it cannot have a [[1,N]][\![1,N]\!]-collision. But for any K[[1,N]]K\subset[\![1,N]\!] with cardinal |K|N1|K|\leq N-1, it holds that

(2) U has a K-collision at t[0,ξ) if and only if X has a K-collision at ρt[0,τ),\displaystyle\hbox{$U$ has a $K$-collision at $t\in[0,\xi)$ if and only if $X$ has a $K$-collision at $\rho_{t}\in[0,\tau)$},

which follows from the facts that

\bullet for all (m,r,u)2×(0,)×𝒰(m,r,u)\in{\mathbb{R}}^{2}\times(0,\infty)\times{\mathcal{U}}, RK(Ψ(m,r,u))=0R_{K}(\Psi(m,r,u))=0 if and only if RK(u)=0R_{K}(u)=0;

\bullet ρ\rho is an increasing bijection from [0,ξ)[0,\xi) into [0,τ)[0,\tau), because ρξ=τ\rho_{\xi}=\tau.

We conclude this subsection with a remark about the quasi-everywhere notions of 𝕏{\mathbb{X}} and 𝕌{\mathbb{U}}, in the case where they are related as above. See Subsection B.1 for a short reminder on this notion.

Remark 9.3.

Fix BUB\in{\mathcal{M}}^{U} such that uU(B)=1{\mathbb{P}}^{U}_{u}(B)=1 quasi-everywhere (here quasi-everywhere refers to the Hunt process 𝕌{\mathbb{U}}). Then Φ𝕊(x)U(B)=1{\mathbb{P}}^{U}_{\Phi_{\mathbb{S}}(x)}(B)=1 quasi-everywhere (here quasi-everywhere refers to the Hunt process 𝕏{\mathbb{X}}^{*}, which is 𝕏{\mathbb{X}} killed when it gets outside ENE_{N}).

Proof.

By definition, there exists 𝒩U{\mathcal{N}}^{U} a properly exceptional set relative to 𝕌{\mathbb{U}} such that for all u𝒰𝒩Uu\in{\mathcal{U}}\setminus{\mathcal{N}}^{U}, uU(B)=1{\mathbb{P}}^{U}_{u}(B)=1. Thus for all xΦ𝕊1(𝒰𝒩U)x\in\Phi_{\mathbb{S}}^{-1}({\mathcal{U}}\setminus{\mathcal{N}}^{U}), Φ𝕊(x)U(B)=1{\mathbb{P}}^{U}_{\Phi_{\mathbb{S}}(x)}(B)=1.

By Proposition 5.1, there exists 𝒩X{\mathcal{N}}^{X} a properly exceptional set relative to 𝕏{\mathbb{X}}^{*}, such that for all x(𝒳EN)𝒩Xx\in({\mathcal{X}}\cap E_{N})\setminus{\mathcal{N}}^{X}, the law of (Xt)t0(X_{t})_{t\geq 0} under xX{\mathbb{P}}^{X}_{x} is equal to the the law of (Yt=Ψ(Mt,Dt,UAt))t0(Y_{t}=\Psi(M_{t},D_{t},U_{A_{t}}))_{t\geq 0} under xY=πH(x)MπH(x)2DΦ𝕊(x)U{\mathbb{Q}}^{Y}_{x}={\mathbb{P}}^{M}_{\pi_{H^{\perp}}(x)}\otimes{\mathbb{P}}^{D}_{\|\pi_{H}(x)\|^{2}}\otimes{\mathbb{P}}^{U}_{\Phi_{\mathbb{S}}(x)}, with some obvious notation.

Hence we only have to prove that 𝒩=Φ𝕊1(𝒩U)𝒩X{\mathcal{N}}=\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U})\cup{\mathcal{N}}^{X} is properly exceptional for 𝕏{\mathbb{X}}^{*}.

\bullet First, we have xX(Xt𝒩{\mathbb{P}}^{X}_{x}(X_{t}^{*}\notin{\mathcal{N}} for all t0)=1t\geq 0)=1 for all x𝒳𝒩x\in{\mathcal{X}}\setminus{\mathcal{N}}. Indeed, since x𝒳𝒩x\in{\mathcal{X}}\setminus{\mathcal{N}}, the law of (Xt)t0(X_{t}^{*})_{t\geq 0} under xX{\mathbb{P}}^{X}_{x} equals the law of (Yt)t0(Y_{t})_{t\geq 0} under xY{\mathbb{Q}}^{Y}_{x}. Since uU(Ut𝒩U{\mathbb{P}}^{U}_{u}(U_{t}\notin{\mathcal{N}}^{U} for all t0)=1t\geq 0)=1 for all u𝒰𝒩Uu\in{\mathcal{U}}\setminus{\mathcal{N}}^{U} and since Φ𝕊(Yt)=UAt\Phi_{\mathbb{S}}(Y_{t})=U_{A_{t}}, we have xY(YtΦ𝕊1(𝒩U){\mathbb{Q}}^{Y}_{x}(Y_{t}\notin\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U}) for all t0)=1t\geq 0)=1 for all x𝒳Φ𝕊1(𝒩U)x\in{\mathcal{X}}\setminus\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U}). Hence xX(XtΦ𝕊1(𝒩U){\mathbb{P}}^{X}_{x}(X_{t}^{*}\notin\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U}) for all t0)=1t\geq 0)=1 for all x𝒳(Φ𝕊1(𝒩U)𝒩X)x\in{\mathcal{X}}\setminus(\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U})\cup{\mathcal{N}}^{X}). Finally, xX(XtΦ𝕊1(𝒩U)𝒩X{\mathbb{P}}^{X}_{x}(X_{t}^{*}\notin\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U})\cup{\mathcal{N}}^{X} for all t0)=1t\geq 0)=1 for all x𝒳(Φ𝕊1(𝒩U)𝒩X)x\in{\mathcal{X}}\setminus(\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U})\cup{\mathcal{N}}^{X}) because 𝒩X{\mathcal{N}}^{X} is properly exceptional for 𝕏{\mathbb{X}}^{*}.

\bullet We have μ(𝒩)=0\mu({\mathcal{N}})=0. Indeed, μ(𝒩X)=0\mu({\mathcal{N}}^{X})=0 by definition and, using Lemma A.2,

μ(Φ𝕊1(𝒩U))=122×+×𝕊1I{Ψ(z,r,u)Φ𝕊1(𝒩U)}rνdzdrβ(du)=122×+β(𝒩U)rνdzdr=0,\displaystyle\mu(\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U}))=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathbb{S}}}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\Psi(z,r,u)\in\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U})\}}r^{\nu}{\rm d}z{\rm d}r\beta({\rm d}u)=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}}\beta({\mathcal{N}}^{U})r^{\nu}{\rm d}z{\rm d}r=0,

because β(𝒩U)=0\beta({\mathcal{N}}^{U})=0. We used that Ψ(z,r,u)Φ𝕊1(𝒩U)u𝒩U\Psi(z,r,u)\in\Phi_{\mathbb{S}}^{-1}({\mathcal{N}}^{U})\Leftrightarrow u\in{\mathcal{N}}^{U}, since Φ𝕊(Ψ(z,r,u))=u\Phi_{\mathbb{S}}(\Psi(z,r,u))=u. ∎

9.2. An expression of dispersion processes on the sphere

We now study the dispersion process (RK(Ut))t0(R_{K}(U_{t}))_{t\geq 0} , for K[[1,N]]K\subset[\![1,N]\!]. The equation below can be informally established if assuming that (1) rigorously holds true, after a time-change and several Itô computations.

Lemma 9.4.

Fix N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta and recall that k0=2N/θk_{0}=\lceil 2N/\theta\rceil. Consider a QSKS(θ,N)QSKS(\theta,N) -process 𝕌{\mathbb{U}} with life-time ξ\xi, fix K[[1,N]]K\subset[\![1,N]\!] such that |K|2|K|\geq 2, and set 𝐊=(K,Kc){\mathbf{K}}=(K,K^{c}). Recall that G𝐊,εG_{{\mathbf{K}},\varepsilon} was introduced in Lemma 6.1, and observe that

G𝐊,0𝕊={u𝒰:miniK,jKuiuj>0}.\color[rgb]{0,0,0}G_{{\mathbf{K}},0}\cap{\mathbb{S}}=\color[rgb]{0,0,0}\Big{\{}u\in{\mathcal{U}}:\min_{i\in K,j\notin K}||u^{i}-u^{j}||>0\Big{\}}.

Quasi-everywhere in G𝐊,0𝕊G_{{\mathbf{K}},0}\cap{\mathbb{S}}, enlarging the filtered probability space (ΩU,U,(tU)t0,uU)(\Omega^{U},{\mathcal{M}}^{U},({\mathcal{M}}^{U}_{t})_{t\geq 0},{\mathbb{P}}^{U}_{u}) if necessary, there exists a 11-dimensional (tU)t0({\mathcal{M}}^{U}_{t})_{t\geq 0}-Brownian motion (Wt)t0(W_{t})_{t\geq 0} under uU{\mathbb{P}}_{u}^{U} such that

(3) RK(Ut)=\displaystyle R_{K}(U_{t})= RK(u)+20tRK(Us)(1RK(Us))dWs+dθ,N(|K|)t\displaystyle R_{K}(u)+2\int_{0}^{t}\sqrt{R_{K}(U_{s})(1-R_{K}(U_{s}))}{\rm d}W_{s}+d_{\theta,N}(|K|)t
dθ,N(N)0tRK(Us)ds2θNiK,jK0tUsiUsjUsiUsj2(UsiSK(Us))ds\displaystyle-d_{\theta,N}(N)\int_{0}^{t}R_{K}(U_{s}){\rm d}s-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\int_{0}^{t}\frac{U_{s}^{i}-U_{s}^{j}}{\|U_{s}^{i}-U_{s}^{j}\|^{2}}\cdot(U_{s}^{i}-S_{K}(U_{s})){\rm d}s

for all t[0,κK)t\in[0,\kappa_{K}), where κK=inf{t0:UtG𝐊,0}\kappa_{K}=\inf\{t\geq 0:U_{t}\notin\color[rgb]{0,0,0}G_{{\mathbf{K}},0}\color[rgb]{0,0,0}\}.

As usual, κKξ\kappa_{K}\leq\xi because G𝐊,0\triangle\notin\color[rgb]{0,0,0}G_{{\mathbf{K}},0}. Note also that if K=[[1,N]]K=[\![1,N]\!], then RK(Ut)=1R_{K}(U_{t})=1 for all t[0,ξ)t\in[0,\xi), and that the constant process 11 indeed solves (3).

Proof.

We divide the proof in several steps. The main idea is to compute URK{\mathcal{L}}^{U}R_{K} and U(RK)2{\mathcal{L}}^{U}(R_{K})^{2} and to use that RK(Ut)=RK(u)+0tURK(Us)ds+MtR_{K}(U_{t})=R_{K}(u)+\int_{0}^{t}{\mathcal{L}}^{U}R_{K}(U_{s}){\rm d}s+M_{t}, for some martingale (Mt)t0(M_{t})_{t\geq 0} of which we can compute the bracket. However, we need to regularize RKR_{K} and to localize space in a zone where the last term of (3) is bounded.

Step 1. We fix n1n\geq 1 and ε(0,1]\varepsilon\in(0,1] and recall Γ𝐊,ε𝕊,nC(𝕊)\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}\in C^{\infty}({\mathbb{S}}), compactly supported in G𝐊,0𝕊G_{{\mathbf{K}},0}\cap{\mathbb{S}}, was defined in Lemma 6.1. We want to apply Remark 4.3 to RKΓ𝐊,ε𝕊,nR_{K}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n} and (RKΓ𝐊,ε𝕊,n)2(R_{K}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n})^{2}. We thus have to show that RKΓ𝐊,ε𝕊,nR_{K}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n} and (RKΓ𝐊,ε𝕊,n)2(R_{K}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n})^{2} belong to Cc(𝒰)C^{\infty}_{c}({\mathcal{U}}) for all n1n\geq 1, which is clear, and that

supα(0,1]supu𝕊(|αU[RKΓ𝐊,ε𝕊,n](u)|+|αU[(RKΓ𝐊,ε𝕊,n)2](u)|)<\sup_{\alpha\in(0,1]}\sup_{u\in{\mathbb{S}}}\Big{(}|{\mathcal{L}}^{U}_{\alpha}[R_{K}\color[rgb]{0,0,0}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}]\color[rgb]{0,0,0}(u)|+|{\mathcal{L}}^{U}_{\alpha}[(R_{K}\color[rgb]{0,0,0}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}\color[rgb]{0,0,0})^{2}](u)|\Big{)}<\infty

for all n1n\geq 1. Since

(4) αU(fg)=fαUg+gαUf+𝕊f𝕊g\displaystyle{\mathcal{L}}^{U}_{\alpha}(fg)=f{\mathcal{L}}^{U}_{\alpha}g+g{\mathcal{L}}^{U}_{\alpha}f+\nabla_{\mathbb{S}}f\cdot\nabla_{\mathbb{S}}g

for all f,gC(𝕊)f,g\in C^{\infty}({\mathbb{S}}) and recalling that supα(0,1]supu𝕊|αUΓ𝐊,ε𝕊,n(u)|<\sup_{\alpha\in(0,1]}\sup_{u\in{\mathbb{S}}}|{\mathcal{L}}^{U}_{\alpha}\color[rgb]{0,0,0}\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n}\color[rgb]{0,0,0}(u)|<\infty by Lemma 6.1 and that Γ𝐊,ε𝕊,n\Gamma_{{\mathbf{K}},\varepsilon}^{{\mathbb{S}},n} is compactly supported in G𝐊,0𝕊G_{{\mathbf{K}},0}\cap{\mathbb{S}}, the only issue is to verify that, for AA compact in G𝐊,0𝕊G_{{\mathbf{K}},0}\cap{\mathbb{S}},

(5) supα(0,1]supuA|αURK(u)|<.\sup_{\alpha\in(0,1]}\sup_{u\in A}|{\mathcal{L}}^{U}_{\alpha}R_{K}(u)|<\infty.

Step 2. Here we prove that

(6) αURK(u)=\displaystyle{\mathcal{L}}^{U}_{\alpha}R_{K}(u)= 2(|K|1)2(N1)RK(u)+θNRK(u)1i,jNuiuj2uiuj2+α\displaystyle 2(|K|-1)-2(N-1)R_{K}(u)+\frac{\theta}{N}R_{K}(u)\sum_{1\leq i,j\leq N}\frac{\|u^{i}-u^{j}\|^{2}}{\|u^{i}-u^{j}\|^{2}+\alpha}
θNiK,jKuiuj2uiuj2+α2θNiK,jKuiujuiuj2+α(uiSK(u)),\displaystyle-\frac{\theta}{N}\sum_{i\in K,j\in K}\frac{\|u^{i}-u^{j}\|^{2}}{\|u^{i}-u^{j}\|^{2}+\alpha}-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(u^{i}-S_{K}(u)),

and this will imply (5): the first four terms are obviously uniformly bounded on 𝕊{\mathbb{S}}, and the last one is uniformly bounded on AA (because AA is compact in G𝐊,0𝕊G_{{\mathbf{K}},0}\cap{\mathbb{S}}). This will also imply, taking α=0\alpha=0 and observing that 2(|K|1)θN|K|(|K|1)=dθ,N(|K|)2(|K|-1)-\frac{\theta}{N}|K|(|K|-1)=d_{\theta,N}(|K|) and 2(N1)θNN(N1)=dθ,N(N)2(N-1)-\frac{\theta}{N}N(N-1)=d_{\theta,N}(N), that for all u𝕊E2u\in{\mathbb{S}}\cap E_{2},

(7) URK(u)=dθ,N(|K|)dθ,N(N)RK(u)2θNiK,jKuiujuiuj2(uiSK(u)).\displaystyle{\mathcal{L}}^{U}R_{K}(u)=d_{\theta,N}(|K|)-d_{\theta,N}(N)R_{K}(u)-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}}\cdot(u^{i}-S_{K}(u)).

Step 2.1. We first verify that for all u𝕊u\in{\mathbb{S}},

(8) (𝕊RK(u))i=2(uiSK(u))1I{iK}2RK(u)ui,i[[1,N]],\displaystyle(\nabla_{\mathbb{S}}R_{K}(u))^{i}=2(u^{i}-S_{K}(u))\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{i\in K\}}-2R_{K}(u)u^{i},\qquad i\in[\![1,N]\!],
(9) Δ𝕊RK(u)=4(|K|1)4(N1)RK(u).\displaystyle\Delta_{\mathbb{S}}R_{K}(u)=4(|K|-1)-4(N-1)R_{K}(u).

First, a simple computation shows that for x(2)Nx\in({\mathbb{R}}^{2})^{N}, for i[[1,N]]i\in[\![1,N]\!],

(10) xiRK(x)=2(xiSK(x))1I{iK}andΔxiRK(x)=4(|K|1)|K|1I{iK},\displaystyle\nabla_{x^{i}}R_{K}(x)=2(x^{i}-S_{K}(x))\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{i\in K\}}\quad\hbox{and}\quad\Delta_{x^{i}}R_{K}(x)=\frac{4(|K|-1)}{|K|}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{i\in K\}},

so that in particular RK(x)H\nabla R_{K}(x)\in H and

(11) RK(x)x=2iK(xiSK(x))xi=2iK(xiSK(x))(xiSK(x))=2RK(x).\displaystyle\nabla R_{K}(x)\cdot x=2\sum_{i\in K}(x^{i}-S_{K}(x))\cdot x^{i}=2\sum_{i\in K}(x^{i}-S_{K}(x))\cdot(x^{i}-S_{K}(x))=2R_{K}(x).

Next, proceeding as in (7), we get [RKΦ𝕊](x)=πH(x)1πH(π(πH(x))(RK(Φ𝕊(x))))\nabla[R_{K}\circ\Phi_{\mathbb{S}}](x)=||\pi_{H}(x)||^{-1}\pi_{H}(\pi_{(\pi_{H}(x))^{\perp}}(\nabla R_{K}(\Phi_{\mathbb{S}}(x)))) for all xENx\in E_{N}, so that

[RKΦ𝕊](x)=πH(RK(Φ𝕊(x))πH(x)RK(Φ𝕊(x))πH(x)2πH(x))πH(x)=RK(x)2RK(x)πH(x)πH(x)2πH(x)2.\nabla[R_{K}\circ\Phi_{\mathbb{S}}](x)=\frac{\pi_{H}\Big{(}\nabla R_{K}(\Phi_{\mathbb{S}}(x))-\frac{\pi_{H}(x)\cdot\nabla R_{K}(\Phi_{\mathbb{S}}(x))}{||\pi_{H}(x)||^{2}}\pi_{H}(x)\Big{)}}{||\pi_{H}(x)||}=\frac{\nabla R_{K}(x)-2R_{K}(x)\frac{\pi_{H}(x)}{||\pi_{H}(x)||^{2}}}{||\pi_{H}(x)||^{2}}.

We used that RK(Φ𝕊(x))=RK(x)/πH(x)\nabla R_{K}(\Phi_{\mathbb{S}}(x))=\nabla R_{K}(x)/\|\pi_{H}(x)\| thanks to (10), that RK(x)H\nabla R_{K}(x)\in H by (10) and that πH(x)RK(x)=xRK(x)=2RK(x)\pi_{H}(x)\cdot\nabla R_{K}(x)=x\cdot\nabla R_{K}(x)=2R_{K}(x) by (11). We first conclude that for u𝕊u\in{\mathbb{S}}, since πH(u)=u\pi_{H}(u)=u and u=1||u||=1,

(12) 𝕊RK(u)=[RKΦ𝕊](u)=RK(u)2RK(u)u,\nabla_{\mathbb{S}}R_{K}(u)=\nabla[R_{K}\circ\Phi_{\mathbb{S}}](u)=\nabla R_{K}(u)-2R_{K}(u)u,

which implies (8) by (10). Second, we deduce that for xENx\in E_{N},

Δ[RKΦ𝕊](x)=\displaystyle\Delta[R_{K}\circ\Phi_{\mathbb{S}}](x)= 1πH(x)2(ΔRK(x)2RK(x)πH(x)πH(x)22RK(x)divπH(x)πH(x)2+4RK(x)πH(x)2)\displaystyle\frac{1}{||\pi_{H}(x)||^{2}}\Big{(}\Delta R_{K}(x)-2\nabla R_{K}(x)\cdot\frac{\pi_{H}(x)}{||\pi_{H}(x)||^{2}}-2R_{K}(x)\frac{{\rm div}\pi_{H}(x)}{||\pi_{H}(x)||^{2}}+\frac{4R_{K}(x)}{||\pi_{H}(x)||^{2}}\Big{)}
2πH(x)πH(x)4(RK(x)2RK(x)πH(x)πH(x)2).\displaystyle-\frac{2\pi_{H}(x)}{||\pi_{H}(x)||^{4}}\cdot\Big{(}\nabla R_{K}(x)-2R_{K}(x)\frac{\pi_{H}(x)}{||\pi_{H}(x)||^{2}}\Big{)}.

Using that divπH(x)=2(N1){\rm div}\,\pi_{H}(x)=2(N-1), we conclude that for u𝕊u\in{\mathbb{S}}, since πH(u)=u\pi_{H}(u)=u, u=1||u||=1 and uRK(u)=2RK(u)u\cdot\nabla R_{K}(u)=2R_{K}(u) by (11),

Δ𝕊RK(u)=Δ[RKΦ𝕊](u)=ΔRK(u)4RK(u)4(N1)RK(u)+4RK(u).\Delta_{\mathbb{S}}R_{K}(u)=\Delta[R_{K}\circ\Phi_{\mathbb{S}}](u)=\Delta R_{K}(u)-4R_{K}(u)-4(N-1)R_{K}(u)+4R_{K}(u).

Since finally ΔRK(u)=4(|K|1)\Delta R_{K}(u)=4(|K|-1) by (10), this leads to (9).

Step 2.2. We fix u𝕊u\in{\mathbb{S}} and show that setting Iα(u)=θN1i,jNuiujuiuj2+α(𝕊RK(u))iI_{\alpha}(u)=-\frac{\theta}{N}\sum_{1\leq i,j\leq N}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(\nabla_{\mathbb{S}}R_{K}(u))^{i}, it holds that

(13) Iα(u)=\displaystyle I_{\alpha}(u)= θNiK,jKuiuj2uiuj2+α+θNRK(u)1i,jNuiuj2uiuj2+α\displaystyle-\frac{\theta}{N}\sum_{i\in K,j\in K}\frac{\|u^{i}-u^{j}\|^{2}}{\|u^{i}-u^{j}\|^{2}+\alpha}+\frac{\theta}{N}R_{K}(u)\sum_{1\leq i,j\leq N}\frac{\|u^{i}-u^{j}\|^{2}}{\|u^{i}-u^{j}\|^{2}+\alpha}
2θNiK,jKuiujuiuj2+α(uiSK(u)).\displaystyle-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(u^{i}-S_{K}(u)).

By (8), we may write Iα=I1,α+I2,αI_{\alpha}=I_{1,\alpha}+I_{2,\alpha}, where

I1,α(u)=2θNiK,j[[1,N]]uiujuiuj2+α(uiSK(u)),\displaystyle I_{1,\alpha}(u)=-\frac{2\theta}{N}\sum_{i\in K,j\in[\![1,N]\!]}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(u^{i}-S_{K}(u)),
I2,α(u)=2θNRK(u)1i,jNuiujuiuj2+αui.\displaystyle I_{2,\alpha}(u)=\frac{2\theta}{N}R_{K}(u)\sum_{1\leq i,j\leq N}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot u^{i}.

First, by symmetry,

I1,α(u)=\displaystyle I_{1,\alpha}(u)= 2θNiK,jKuiujuiuj2+α(uiSK(u))2θNiK,jKuiujuiuj2+α(uiSK(u))\displaystyle-\frac{2\theta}{N}\sum_{i\in K,j\in K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(u^{i}-S_{K}(u))-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(u^{i}-S_{K}(u))
=\displaystyle= 2θNiK,jKuiujuiuj2+αui2θNiK,jKuiujuiuj2+α(uiSK(u))\displaystyle-\frac{2\theta}{N}\sum_{i\in K,j\in K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot u^{i}-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(u^{i}-S_{K}(u))
=\displaystyle= θNiK,jKuiuj2uiuj2+α2θNiK,jKuiujuiuj2+α(uiSK(u)).\displaystyle-\frac{\theta}{N}\sum_{i\in K,j\in K}\frac{\|u^{i}-u^{j}\|^{2}}{\|u^{i}-u^{j}\|^{2}+\alpha}-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{u^{i}-u^{j}}{\|u^{i}-u^{j}\|^{2}+\alpha}\cdot(u^{i}-S_{K}(u)).

Second, by symmetry,

I2,α(u)=θNRK(u)1i,jNuiuj2uiuj2+α.I_{2,\alpha}(u)=\frac{\theta}{N}R_{K}(u)\sum_{1\leq i,j\leq N}\frac{\|u^{i}-u^{j}\|^{2}}{\|u^{i}-u^{j}\|^{2}+\alpha}.

Step 2.3. Since αURK(u)=12Δ𝕊RK(u)+Iα(u){\mathcal{L}}^{U}_{\alpha}R_{K}(u)=\frac{1}{2}\Delta_{\mathbb{S}}R_{K}(u)+I_{\alpha}(u), (6) follows from (9) and (13).

Step 3. By Steps 1 and 2, we can apply Remark 4.3 and Lemma B.2: quasi-everywhere, for all n1n\geq 1, there exist two (tU)t0({\mathcal{M}}^{U}_{t})_{t\geq 0}-martingales (Mt1,n,ε)t0(M^{1,n,\varepsilon}_{t})_{t\geq 0} and (Mt2,n,ε)t0(M^{2,n,\varepsilon}_{t})_{t\geq 0} under uU{\mathbb{P}}^{U}_{u}, such that

(RKΓ𝐊,ε𝕊,n)(Ut)=(RKΓ𝐊,ε𝕊,n)(u)+Mt1,n,ε+0tU(RKΓ𝐊,ε𝕊,n)(Us)ds,\displaystyle\color[rgb]{0,0,0}(R_{K}\Gamma^{{\mathbb{S}},n}_{{\mathbf{K}},\varepsilon})(U_{t})=(R_{K}\color[rgb]{0,0,0}\Gamma^{{\mathbb{S}},n}_{{\mathbf{K}},\varepsilon})(u)+M^{1,n,\varepsilon}_{t}+\int_{0}^{t}{\mathcal{L}}^{U}(R_{K}\Gamma^{{\mathbb{S}},n}_{{\mathbf{K}},\varepsilon})(U_{s}){\rm d}s,
(RKΓ𝐊,ε𝕊,n)2(Ut)=(RKΓ𝐊,ε𝕊,n)2(u)+Mt2,n,ε+0tU(RKΓ𝐊,ε𝕊,n)2(Us)ds\displaystyle(R_{K}\Gamma^{{\mathbb{S}},n}_{{\mathbf{K}},\varepsilon})^{2}(U_{t})=(R_{K}\Gamma^{{\mathbb{S}},n}_{{\mathbf{K}},\varepsilon})^{2}(u)+M^{2,n,\varepsilon}_{t}+\int_{0}^{t}{\mathcal{L}}^{U}(R_{K}\Gamma^{{\mathbb{S}},n}_{{\mathbf{K}},\varepsilon})^{2}(U_{s}){\rm d}s

for all t0t\geq 0. We recall that κK=inf{t0:UtG𝐊,0n}\kappa_{K}=\inf\{t\geq 0:U_{t}\notin G_{{\mathbf{K}},0}^{n}\} and introduce

κK,n,ε=inf{t0:UtG𝐊,εn}κK.\kappa_{K,n,\varepsilon}=\inf\{t\geq 0:U_{t}\notin\color[rgb]{0,0,0}G_{{\mathbf{K}},\varepsilon}^{n}\}\land\kappa_{K}.

Since n1G𝐊,εnG𝐊,ε\cup_{n\geq 1}G_{{\mathbf{K}},\varepsilon}^{n}\supset G_{{\mathbf{K}},\varepsilon} and since G𝐊,εG_{{\mathbf{K}},\varepsilon} increases to G𝐊,0G_{{\mathbf{K}},0} as ε0\varepsilon\to 0, see Lemma 6.1, we conclude that limε0limnκK,n,ε=κK\lim_{\varepsilon\to 0}\lim_{n\to\infty}\kappa_{K,n,\varepsilon}=\kappa_{K}. Next, since Γ𝐊,ε𝕊,n=1\Gamma^{{\mathbb{S}},n}_{{\mathbf{K}},\varepsilon}=1 on G𝐊,εn𝕊G_{{\mathbf{K}},\varepsilon}^{n}\cap{\mathbb{S}}, we have, for all t[0,κK,n,ε]t\in[0,\kappa_{K,n,\varepsilon}],

(14) RK(Ut)=RK(u)+Mt1,n,ε+0tURK(Us)ds,\displaystyle R_{K}(U_{t})=R_{K}(u)+M^{1,n,\varepsilon}_{t}+\int_{0}^{t}{\mathcal{L}}^{U}R_{K}(U_{s}){\rm d}s,
(15) (RK(Ut))2=(RK(u))2+Mt2,n,ε+0tU(RK2)(Us)ds.\displaystyle(R_{K}(U_{t}))^{2}=(R_{K}(u))^{2}+M^{2,n,\varepsilon}_{t}+\int_{0}^{t}{\mathcal{L}}^{U}(R_{K}^{2})(U_{s}){\rm d}s.

Applying the Itô formula to compute (RK(Ut))2(R_{K}(U_{t}))^{2} from (14), recalling from (4) that U(RK2)=2RKURK+𝕊RK2{\mathcal{L}}^{U}(R_{K}^{2})=2R_{K}{\mathcal{L}}^{U}R_{K}+||\nabla_{\mathbb{S}}R_{K}||^{2} and comparing to (15), we obtain that for t[0,κK,n,ε]t\in[0,\kappa_{K,n,\varepsilon}],

M1,n,εt=0t𝕊RK(Us)2ds.\langle M^{1,n,\varepsilon}\rangle_{t}=\int_{0}^{t}\|\nabla_{\mathbb{S}}R_{K}(U_{s})\|^{2}{\rm d}s.

Hence, enlarging the probability space if necessary, we can find a Brownian motion (Wt)t0(W_{t})_{t\geq 0}, which is defined by Wt=0t𝕊RK(Us)1dMs1,n,εW_{t}=\int_{0}^{t}\|\nabla_{\mathbb{S}}R_{K}(U_{s})\|^{-1}{\rm d}M^{1,n,\varepsilon}_{s} for t[0,κK,n,ε]t\in[0,\kappa_{K,n,\varepsilon}] and which is then extended to +{\mathbb{R}}_{+}, such that Mt1,n,ε=0t𝕊RK(Us)dWsM^{1,n,\varepsilon}_{t}=\int_{0}^{t}\|\nabla_{\mathbb{S}}R_{K}(U_{s})\|{\rm d}W_{s} during [0,κK,n,ε][0,\kappa_{K,n,\varepsilon}]. Hence, still for t[0,κK,n,ε]t\in[0,\kappa_{K,n,\varepsilon}],

(16) RK(Ut)=RK(u)+0t𝕊RK(Us)dWs+0tURK(Us)ds.R_{K}(U_{t})=R_{K}(u)+\int_{0}^{t}\|\nabla_{\mathbb{S}}R_{K}(U_{s})\|{\rm d}W_{s}+\int_{0}^{t}{\mathcal{L}}^{U}R_{K}(U_{s}){\rm d}s.

But 𝕊RK(u)=RK(u)2RK(u)u\nabla_{\mathbb{S}}R_{K}(u)=\nabla R_{K}(u)-2R_{K}(u)u by (12), whence

𝕊RK(u)2=RK(u)24RK(u)RK(u)u+4(RK(u))2.\|\nabla_{\mathbb{S}}R_{K}(u)\|^{2}=\|\nabla R_{K}(u)\|^{2}-4R_{K}(u)\nabla R_{K}(u)\cdot u+4(R_{K}(u))^{2}.

Since RK(u)2=4RK(u)||\nabla R_{K}(u)||^{2}=4R_{K}(u) by (10) and RK(u)u=2RK(u)\nabla R_{K}(u)\cdot u=2R_{K}(u) by (11),

𝕊RK(u)2=4RK(u)4(RK(u))2=4RK(u)(1RK(u)).\|\nabla_{\mathbb{S}}R_{K}(u)\|^{2}=4R_{K}(u)-4(R_{K}(u))^{2}=4R_{K}(u)(1-R_{K}(u)).

Inserting this, as well as the expression (7) of URK{\mathcal{L}}^{U}R_{K}, in (16), shows that RK(Ut)R_{K}(U_{t}) satisfies the desired equation on [0,κK,n,ε][0,\kappa_{K,n,\varepsilon}]. Since limε0limnκK,n,ε=κK\lim_{\varepsilon\to 0}\lim_{n\to\infty}\kappa_{K,n,\varepsilon}=\kappa_{K} a.s., the proof is complete. ∎

9.3. A squared Bessel-like process

The equation obtained in the previous lemma will be studied by comparison with the process we now introduce. This process behaves, near 0, like a squared Bessel processes.

Lemma 9.5.

Fix δ\delta\in{\mathbb{R}}, a>0a>0 and b>0b>0 such that δ+ab<2\delta+a\sqrt{b}<2. For (Wt)t0(W_{t})_{t\geq 0} a 11-dimensional Brownian motion and for x[0,1)x\in[0,1), consider the unique solution (St)t0(S_{t})_{t\geq 0} of

(17) St=x+0t2|Ss(1Ss)|dWs+δt+a0tb+|Ss|ds.\displaystyle S_{t}=x+\int_{0}^{t}2\sqrt{|S_{s}(1-S_{s})|}{\rm d}W_{s}+\delta t+a\int_{0}^{t}\sqrt{b+|S_{s}|}{\rm d}s.

For zz\in{\mathbb{R}}, set τz=inf{t>0:St=z}\tau_{z}=\inf\{t>0:S_{t}=z\}. For all y(x,1)y\in(x,1), it holds that (τ0<τy)>0{\mathbb{P}}(\tau_{0}<\tau_{y})>0.

Proof.

This equation is classically well-posed, since the diffusion coefficient is 1/21/2-Hölder continuous and the drift coefficient is Lipschitz continuous, see Revuz-Yor [21, Theorem 3.5 page 390]. As in Karatzas-Shreve [15, (5.42) page 339], we introduce the scale function

f(z)=1/2zexp(1/2uδ+ab+|v|2|v(1v)|dv)du.f(z)=\int_{1/2}^{z}\exp\Big{(}-\int_{1/2}^{u}\frac{\delta+a\sqrt{b+|v|}}{2|v(1-v)|}{\rm d}v\Big{)}{\rm d}u.

This function is obviously continuous on (0,1)(0,1) and one gets convinced, for example approximating (δ+ab+|v|)/(2|v(1v)|)(\delta+a\sqrt{b+|v|})/(2|v(1-v)|) by (δ+ab)/(2|v|)(\delta+a\sqrt{b})/(2|v|), that it is also continuous at 0 because δ+ab<2\delta+a\sqrt{b}<2. By [15, (5.61) page 344], we have

(18) (τ0<τy)=f(y)f(x)f(y)f(0).{\mathbb{P}}(\tau_{0}<\tau_{y})=\frac{f(y)-f(x)}{f(y)-f(0)}.

for all y(x,1)y\in(x,1). This last quantity is nonzero (which would not be the case if δ+ab2\delta+a\sqrt{b}\geq 2, since then f(0)=f(0)=-\infty). ∎

9.4. Collisions of large clusters

We are now ready to give the

Proof of Proposition 9.1-(i)-(ii).

We fix N4N\geq 4, θ>0\theta>0 such that N>θN>\theta. We always assume that dθ,N(N)<2d_{\theta,N}(N)<2 and we use the notation of Subsection 9.1. Step 1. We consider ε(0,1]\varepsilon\in(0,1] and K[[1,N]]K\subset[\![1,N]\!] such that |K|[[2,N1]]|K|\in[\![2,N-1]\!] and dθ,N(|K|)<2d_{\theta,N}(|K|)<2. We introduce the constant aK=c|K|+1/(2c|K|)a_{K}=c_{|K|+1}/(2c_{|K|}) with (c)[[1,N]](c_{\ell})_{\ell\in[\![1,N]\!]} defined in Lemma 6.2. We prove in this step that there are some constants pK,ε>0p_{K,\varepsilon}>0 and TK,ε>0T_{K,\varepsilon}>0 such that, setting

σ~K,ε=inf{t>0:RK(Ut)ε or miniKRK{i}(Ut)aKε}TK,ε,\displaystyle\tilde{\sigma}^{K,\varepsilon}=\inf\Big{\{}t>0:R_{K}(U_{t})\geq\varepsilon\;\mbox{ or }\;\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\leq a_{K}\varepsilon\Big{\}}\land T_{K,\varepsilon},

with the convention that inf=ξ\inf\emptyset=\xi, it holds that quasi-everywhere on {u𝒰:RK(u)ε/2}\{u\in{\mathcal{U}}:R_{K}(u)\leq\varepsilon/2\},

uU(σ~K,ε=ξ or inft[0,σ~K,ε)miniKRK{i}(Ut)2aKε or RK(Ut)=0 for some t[0,σ~K,ε))pK,ε.{\mathbb{P}}^{U}_{u}\Big{(}\tilde{\sigma}^{K,\varepsilon}=\xi\;\mbox{ or }\inf_{t\in[0,\tilde{\sigma}^{K,\varepsilon})}\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\leq 2a_{K}\varepsilon\hbox{ or }R_{K}(U_{t})=0\mbox{ for some }t\in[0,\tilde{\sigma}^{K,\varepsilon})\Big{)}\geq p_{K,\varepsilon}.

We introduce ZK,ε=inft[0,σ~K,ε)miniKRK{i}(Ut)Z_{K,\varepsilon}=\inf_{t\in[0,\tilde{\sigma}^{K,\varepsilon})}\min_{i\notin K}R_{K\cup\{i\}}(U_{t}). We note that for all t[0,σ~K,ε)t\in[0,\tilde{\sigma}^{K,\varepsilon}), RK(Ut)εR_{K}(U_{t})\leq\varepsilon and ZK,εaKε\color[rgb]{0,0,0}Z_{K,\varepsilon}\color[rgb]{0,0,0}\geq a_{K}\varepsilon so that miniK,jKUtiUtjε/2\min_{i\in K,j\notin K}\|U^{i}_{t}-U^{j}_{t}\|\geq\varepsilon/2 thanks to the definition of aKa_{K} and to Lemma 6.2. This implies that σ~K,εκK\tilde{\sigma}^{K,\varepsilon}\leq\kappa_{K}, where we recall that κK=inf{t0:UtG𝐊,0}\kappa_{K}=\inf\{t\geq 0:U_{t}\notin\color[rgb]{0,0,0}G_{{\mathbf{K}},0}\color[rgb]{0,0,0}\} was defined in Lemma 9.4, and that G𝐊,0𝕊={u𝒰:miniK,jKuiuj>0}\color[rgb]{0,0,0}G_{{\mathbf{K}},0}\cap{\mathbb{S}}\color[rgb]{0,0,0}=\{u\in{\mathcal{U}}:\min_{i\in K,j\notin K}||u^{i}-u^{j}||>0\}.

By the Cauchy-Schwarz inequality, and since RKR_{K} is bounded on 𝒰{\mathcal{U}}, there is a deterministic constant CK,ε>0C_{K,\varepsilon}>0, allowed to change from line to line, such that for all t[0,σ~K,ε)t\in[0,\tilde{\sigma}^{K,\varepsilon}), we have

dθ,N(N)RK(Ut)2θNiK,jKUtiUtjUtiUtj2(UtiSK(Ut))\displaystyle-d_{\theta,N}(N)R_{K}(U_{t})-\frac{2\theta}{N}\sum_{i\in K,j\notin K}\frac{U_{t}^{i}-U_{t}^{j}}{\|U_{t}^{i}-U_{t}^{j}\|^{2}}\cdot(U_{t}^{i}-S_{K}(U_{t}))
\displaystyle\leq CK,εRK(Ut)+CK,ε(iKUtiSK(Ut)2)1/2\displaystyle C_{K,\varepsilon}\sqrt{R_{K}(U_{t})}+C_{K,\varepsilon}\Big{(}\sum_{i\in K}\|U_{t}^{i}-S_{K}(U_{t})\|^{2}\Big{)}^{1/2}
\displaystyle\leq CK,εRK(Ut)\displaystyle C_{K,\varepsilon}\sqrt{R_{K}(U_{t})}
\displaystyle\leq CK,εb+RK(Ut)\displaystyle C_{K,\varepsilon}\sqrt{b+R_{K}(U_{t})}

where b>0b>0 is chosen small enough so that dθ,N(|K|)+CK,εb<2d_{\theta,N}(|K|)+C_{K,\varepsilon}\sqrt{b}<2. Actually, bb is only introduced to make the drift coefficient of (17) Lipschitz continuous.

Recalling that RK(U0)ε/2R_{K}(U_{0})\leq\varepsilon/2, the formula describing RK(Ut)[0,1]R_{K}(U_{t})\in[0,1] for t[0,κK)[0,σ~K,ε)t\in[0,\kappa_{K})\supset[0,\tilde{\sigma}^{K,\varepsilon}), see Lemma 9.4, considering the process (St)t0(S_{t})_{t\geq 0} solution to (17) with x=ε/2x=\varepsilon/2, δ=dθ,N(|K|)\delta=d_{\theta,N}(|K|), a=CK,εa=C_{K,\varepsilon} and with bb introduced a few lines above, driven by the same Brownian motion (Wt)t0(W_{t})_{t\geq 0}, and using the comparison theorem, we conclude that RK(Ut)StR_{K}(U_{t})\leq S_{t} for all t[0,σ~K,ε)t\in[0,\tilde{\sigma}^{K,\varepsilon}).

Setting τz=inf{t0:St=z}\tau_{z}=\inf\{t\geq 0:S_{t}=z\} for zz\in{\mathbb{R}} and recalling the definition of σ~K,ε\tilde{\sigma}^{K,\varepsilon}, we conclude that {ZK,ε>2aKε}{σ~K,ετεTK,ε}.\{Z_{K,\varepsilon}>2a_{K}\varepsilon\}\subset\{\tilde{\sigma}^{K,\varepsilon}\geq\tau_{\varepsilon}\land T_{K,\varepsilon}\}. Indeed, on {inft[0,σ~K,ε)miniKRK{i}(Ut)>2aKε}\{\inf_{t\in[0,\tilde{\sigma}^{K,\varepsilon})}\min_{i\notin K}R_{K\cup\{i\}}(U_{t})>2a_{K}\varepsilon\}, either σ~K,ε=TK,ε\tilde{\sigma}^{K,\varepsilon}=T_{K,\varepsilon}, or (RK(Ut))t0(R_{K}(U_{t}))_{t\geq 0} reaches ε\varepsilon at time σ~K,ε\tilde{\sigma}^{K,\varepsilon} and we then have τεσ~K,ε\tau_{\varepsilon}\leq\tilde{\sigma}^{K,\varepsilon}. In both cases, σ~K,ετεTK,ε\tilde{\sigma}^{K,\varepsilon}\geq\tau_{\varepsilon}\land T_{K,\varepsilon}. Hence, using again that RK(Ut)StR_{K}(U_{t})\leq S_{t} for all t[0,σ~K,ε)t\in[0,\tilde{\sigma}^{K,\varepsilon}),

{σ~K,ε<ξ and ZK,ε>2aKε and St=0 for some t[0,τεTK,ε]}\displaystyle\Big{\{}\tilde{\sigma}^{K,\varepsilon}<\xi\;\hbox{ and }\;\color[rgb]{0,0,0}Z_{K,\varepsilon}\color[rgb]{0,0,0}>2a_{K}\varepsilon\;\mbox{ and }\;S_{t}=0\mbox{ for some }t\in[0,\tau_{\varepsilon}\land T_{K,\varepsilon}]\Big{\}}
\displaystyle\subset {σ~K,ε<ξ and ZK,ε>2aKε and RK(Ut)=0 for some t[0,σ~K,ε)}.\displaystyle\Big{\{}\tilde{\sigma}^{K,\varepsilon}<\xi\;\mbox{ and }\;\color[rgb]{0,0,0}Z_{K,\varepsilon}\color[rgb]{0,0,0}>2a_{K}\varepsilon\;\hbox{ and }\;R_{K}(U_{t})=0\mbox{ for some }t\in[0,\tilde{\sigma}^{K,\varepsilon})\Big{\}}.

But AcBAcBA^{c}\cap B^{\prime}\subset A^{c}\cap B gives (AB)=(A)+(AcB)(A)+(AcB)=(AB){\mathbb{P}}(A\cup B)={\mathbb{P}}(A)+{\mathbb{P}}(A^{c}\cap B)\geq{\mathbb{P}}(A)+{\mathbb{P}}(A^{c}\cap B^{\prime})={\mathbb{P}}(A\cup B^{\prime}). Hence

uU(σ~K,ε=ξ or ZK,ε2aKε or RK(Ut)=0 for some t[0,σ~K,ε))\displaystyle{\mathbb{P}}^{U}_{u}\Big{(}\tilde{\sigma}^{K,\varepsilon}=\xi\;\mbox{ or }\;\color[rgb]{0,0,0}Z_{K,\varepsilon}\color[rgb]{0,0,0}\leq 2a_{K}\varepsilon\;\hbox{ or }\;R_{K}(U_{t})=0\mbox{ for some }t\in[0,\tilde{\sigma}^{K,\varepsilon})\Big{)}
\displaystyle\geq uU(σ~K,ε=ξ or ZK,ε2aKε or St=0 for some t[0,τεTK,ε))\displaystyle{\mathbb{P}}^{U}_{u}\Big{(}\tilde{\sigma}^{K,\varepsilon}=\xi\;\hbox{ or }\;\color[rgb]{0,0,0}Z_{K,\varepsilon}\color[rgb]{0,0,0}\leq 2a_{K}\varepsilon\;\mbox{ or }\;S_{t}=0\mbox{ for some }t\in[0,\tau_{\varepsilon}\land T_{K,\varepsilon})\Big{)}
\displaystyle\geq uU(St=0 for some t[0,τεTK,ε)).\displaystyle{\mathbb{P}}^{U}_{u}\Big{(}S_{t}=0\mbox{ for some }t\in[0,\tau_{\varepsilon}\land T_{K,\varepsilon})\Big{)}.

This last quantity equals (τ0<τεTK,ε){\mathbb{P}}(\tau_{0}<\tau_{\varepsilon}\land T_{K,\varepsilon}) and does not depend on uu such that RK(u)ε/2R_{K}(u)\leq\varepsilon/2. But (τ0<τε)>0{\mathbb{P}}(\tau_{0}<\tau_{\varepsilon})>0 by Lemma 9.5 and since dθ,N(|K|)+CK,εb<2d_{\theta,N}(|K|)+C_{K,\varepsilon}\sqrt{b}<2. Hence there exists TK,ε>0T_{K,\varepsilon}>0 so that (τ0<τεTK,ε)>0{\mathbb{P}}(\tau_{0}<\tau_{\varepsilon}\land T_{K,\varepsilon})>0 and this completes the step.

Step 2. We prove (ii), i.e. that when dθ,N(N1)(0,2)d_{\theta,N}(N-1)\in(0,2), for any K[[1,N]]K\subset[\![1,N]\!] with cardinal |K|=N1|K|=N-1, quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., RK(Xt)R_{K}(X_{t}) vanishes during [0,ζ)[0,\zeta). By (2) and Remark 9.3, and since uU(ξ=)=1{\mathbb{P}}^{U}_{u}(\xi=\infty)=1 quasi-everywhere by Lemma 4.4-(ii), it suffices to check that quasi-everywhere, uU{\mathbb{P}}^{U}_{u}-a.s., (RK(Ut))t0(R_{K}(U_{t}))_{t\geq 0} vanishes at least once during [0,)[0,\infty).

We fix K[[1,N]]K\subset[\![1,N]\!] with |K|=N1|K|=N-1, set ε0=1/(4aK)\varepsilon_{0}=1/(4a_{K}) and introduce τ~0K=0\tilde{\tau}^{K}_{0}=0 and for all k0k\geq 0,

τk+1K=inf{tτ~kK:RK(Ut)ε0/2},\displaystyle\tau^{K}_{k+1}=\inf\{t\geq\tilde{\tau}^{K}_{k}:R_{K}(U_{t})\leq\varepsilon_{0}/2\},
τ~k+1K=inf{tτk+1K:RK(Ut)ε0}(τk+1K+TK,ε0).\displaystyle\tilde{\tau}^{K}_{k+1}=\inf\{t\geq\tau^{K}_{k+1}:R_{K}(U_{t})\geq\varepsilon_{0}\}\land(\tau^{K}_{k+1}+T_{K,\varepsilon_{0}}).

with TK,ε0T_{K,\varepsilon_{0}} defined in Step 1. All these stopping times are finite since (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is recurrent by Lemma 4.4-(ii). We also put, for k1k\geq 1,

ΩkK={RK(Ut)=0 for some t[τkK,τ~kK]}.\Omega_{k}^{K}=\{R_{K}(U_{t})=0\hbox{ for some }t\in[\tau^{K}_{k},\tilde{\tau}^{K}_{k}]\}.

We now prove that uU(k1(ΩkK)c)=0{\mathbb{P}}^{U}_{u}(\cap_{k\geq 1}(\Omega_{k}^{K})^{c})=0 quasi-everywhere, and this will complete the proof of (ii). For 1\ell\geq 1, since k=1(ΩkK)c\cap_{k=1}^{\ell}(\Omega_{k}^{K})^{c} is τ+1KU{\mathcal{M}}^{U}_{\tau^{K}_{\ell+1}}-measurable, the strong Markov property tells us that

uU(k=1+1(ΩkK)c)=𝔼uU[(k=11I(ΩkK)c)Uτ+1KU((Ω1K)c)].{\mathbb{P}}^{U}_{u}\Big{(}\cap_{k=1}^{\ell+1}(\Omega_{k}^{K})^{c}\Big{)}={\mathbb{E}}^{U}_{u}\Big{[}\Big{(}\prod_{k=1}^{\ell}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{(\Omega_{k}^{K})^{c}}\Big{)}{\mathbb{P}}^{U}_{U_{\tau^{K}_{\ell+1}}}((\Omega_{1}^{K})^{c})\Big{]}.

We now prove that uU(Ω1K)pK,ε0{\mathbb{P}}^{U}_{u}(\Omega_{1}^{K})\geq p_{K,\varepsilon_{0}} quasi-everywhere on {u𝒰:RK(u)ε0/2}\{u\in{\mathcal{U}}:R_{K}(u)\leq\varepsilon_{0}/2\}. For such a uu, we have τ1K=0\tau^{K}_{1}=0. Moreover, for all iKi\notin K, we have RK{i}(u)=R[[1,N]](u)=1>2aKε0R_{K\cup\{i\}}(u)=R_{[\![1,N]\!]}(u)=1>2a_{K}\varepsilon_{0} thanks to our choice of ε0\varepsilon_{0}. Hence τ~1K=σ~K,ε0\tilde{\tau}^{K}_{1}=\tilde{\sigma}^{K,\varepsilon_{0}}, recall Step 1. Since finally σ~K,ε0<=ξ\tilde{\sigma}^{K,\varepsilon_{0}}<\infty=\xi and since RK{i}(Ut)=R[[1,N]](Ut)=1>2aKε0R_{K\cup\{i\}}(U_{t})=R_{[\![1,N]\!]}(U_{t})=1>2a_{K}\varepsilon_{0} for all t0t\geq 0 and all iKi\notin K,

Ω1K=\displaystyle\Omega_{1}^{K}= {RK(Ut)=0 for some t[0,σ~K,ε0]}\displaystyle\Big{\{}R_{K}(U_{t})=0\hbox{ for some }t\in[0,\tilde{\sigma}^{K,\varepsilon_{0}}]\Big{\}}
=\displaystyle= {σ~K,ε0=ξ or inft[0,σ~K,ε0)miniKRK{i}(Ut)2aKε0 or RK(Ut)=0 for some t[0,σ~K,ε0]}.\displaystyle\Big{\{}\tilde{\sigma}^{K,\varepsilon_{0}}=\xi\;\hbox{ or }\inf_{t\in[0,\tilde{\sigma}^{K,\varepsilon_{0}})}\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\leq 2a_{K}\varepsilon_{0}\;\hbox{ or }\;R_{K}(U_{t})=0\hbox{ for some }t\in[0,\tilde{\sigma}^{K,\varepsilon_{0}}]\Big{\}}.

Hence Step 1 tells us that uU(Ω1K)pK,ε0{\mathbb{P}}^{U}_{u}(\Omega_{1}^{K})\geq p_{K,\varepsilon_{0}} quasi-everywhere on {u𝒰:RK(u)ε0/2}\{u\in{\mathcal{U}}:R_{K}(u)\leq\varepsilon_{0}/2\}.

Since RK(Uτ+1K)ε0/2R_{K}(U_{\tau^{K}_{\ell+1}})\leq\varepsilon_{0}/2, we have proved that for all 1\ell\geq 1,

uU(k=1+1(ΩkK)c)(1pK,ε0)uU(k=1(ΩkK)c).{\mathbb{P}}^{U}_{u}\Big{(}\cap_{k=1}^{\ell+1}(\Omega_{k}^{K})^{c}\Big{)}\leq(1-p_{K,\varepsilon_{0}}){\mathbb{P}}^{U}_{u}\Big{(}\cap_{k=1}^{\ell}(\Omega_{k}^{K})^{c}\Big{)}.

This allows us to conclude that indeed, uU(k=1(ΩkK)c)=0{\mathbb{P}}^{U}_{u}(\cap_{k=1}^{\infty}(\Omega_{k}^{K})^{c})=0.

Step 3. We prove (i), i.e. that if dθ,N(N1)0d_{\theta,N}(N-1)\leq 0, then xX(inf[0,ζ)R[[1,N]](Xt)>0)=1{\mathbb{P}}^{X}_{x}(\inf_{[0,\zeta)}R_{[\![1,N]\!]}(X_{t})>0)=1 quasi-everywhere. By Remark 9.3 and (1), it suffices to show that quasi-everywhere, uU(ξ<)=1{\mathbb{P}}^{U}_{u}(\xi<\infty)=1.

For all K[[1,N]]K\subset[\![1,N]\!], all ε(0,1]\varepsilon\in(0,1], we introduce σ~0K,ε=0\tilde{\sigma}^{K,\varepsilon}_{0}=0 and for all k0k\geq 0,

σk+1K,ε=inf{tσ~kK,ε:RK(Ut)ε/2 and miniKRK{i}(Ut)2aKε},\displaystyle\sigma^{K,\varepsilon}_{k+1}=\inf\Big{\{}t\geq\tilde{\sigma}^{K,\varepsilon}_{k}:R_{K}(U_{t})\leq\varepsilon/2\;\mbox{ and }\;\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\geq 2a_{K}\varepsilon\Big{\}},
σ~k+1K,ε=inf{tσk+1K,ε:RK(Ut)ε or miniKRK{i}(Ut)aKε}(σk+1K,ε+TK,ε),\displaystyle\tilde{\sigma}_{k+1}^{K,\varepsilon}=\inf\Big{\{}t\geq\sigma_{k+1}^{K,\varepsilon}:R_{K}(U_{t})\geq\varepsilon\;\mbox{ or }\;\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\leq a_{K}\varepsilon\Big{\}}\land(\sigma^{K,\varepsilon}_{k+1}+T_{K,\varepsilon}),

with TK,εT_{K,\varepsilon} defined in Step 1 and with the convention that inf=ξ\inf\emptyset=\xi. Step 3.1. We fix ε(0,1]\varepsilon\in(0,1] and assume that |K|k0|K|\geq k_{0}, so that dθ,N(|K|)0d_{\theta,N}(|K|)\leq 0 by Lemma 1.1. We prove here that quasi-everywhere, uU{\mathbb{P}}^{U}_{u}-a.s., either there is t[0,ξ)t\in[0,\xi) such that RK(Ut)=0R_{K}(U_{t})=0 or there is k1k\geq 1 such that σk+1K,ε=ξ\sigma^{K,\varepsilon}_{k+1}=\xi or there is k1k\geq 1 such that inft[σkK,ε,σ~kK,ε)miniKRK{i}(Ut)2aKε\inf_{t\in[\sigma^{K,\varepsilon}_{k},\tilde{\sigma}^{K,\varepsilon}_{k})}\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\leq 2a_{K}\varepsilon. It suffices to prove that uU(k1(ΩkK,ε)c)=0{\mathbb{P}}^{U}_{u}(\cap_{k\geq 1}(\Omega_{k}^{K,\varepsilon})^{c})=0, where

ΩkK,ε={σk+1K,ε=ξ or inft[σkK,ε,σ~kK,ε)miniKRK{i}(Ut)2aKε\displaystyle\Omega^{K,\varepsilon}_{k}=\Big{\{}\sigma^{K,\varepsilon}_{k+1}=\xi\,\mbox{ or }\inf_{t\in[\sigma^{K,\varepsilon}_{k},\tilde{\sigma}^{K,\varepsilon}_{k})}\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\leq 2a_{K}\varepsilon
or RK(Ut)=0 for some t[σkK,ε,σ~kK,ε)}.\displaystyle\hskip 142.26378pt\mbox{ or }\,R_{K}(U_{t})=0\hbox{ for some }t\in[\sigma^{K,\varepsilon}_{k},\tilde{\sigma}^{K,\varepsilon}_{k})\Big{\}}.

But for all 1\ell\geq 1, k=1(ΩkK,ε)c\cap_{k=1}^{\ell}(\Omega_{k}^{K,\varepsilon})^{c} is σ+1K,εU{\mathcal{M}}^{U}_{\sigma^{K,\varepsilon}_{\ell+1}}-measurable, whence, by the strong Markov property,

uU(k=1+1(ΩkK,ε)c)=𝔼uU[(k=11I(ΩkK,ε)c)Uσ+1K,εU((Ω1K,ε)c)](1pK,ε)uU(k=1(ΩkK,ε)c).{\mathbb{P}}^{U}_{u}\Big{(}\cap_{k=1}^{\ell+1}(\Omega_{k}^{K,\varepsilon})^{c}\Big{)}={\mathbb{E}}^{U}_{u}\Big{[}\Big{(}\prod_{k=1}^{\ell}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{(\Omega_{k}^{K,\varepsilon})^{c}}\Big{)}{\mathbb{P}}^{U}_{U_{\sigma^{K,\varepsilon}_{\ell+1}}}\Big{(}(\Omega_{1}^{K,\varepsilon})^{c}\Big{)}\Big{]}\leq(1-p_{K,\varepsilon}){\mathbb{P}}^{U}_{u}\Big{(}\cap_{k=1}^{\ell}(\Omega_{k}^{K,\varepsilon})^{c}\Big{)}.

We used Step 1, that RK(Uσ+1K,ε)ε/2R_{K}(U_{\sigma^{K,\varepsilon}_{\ell+1}})\leq\varepsilon/2 on the event (ΩK,ε)c{σ+1K,ε<ξ}(\Omega_{\ell}^{K,\varepsilon})^{c}\subset\{\sigma^{K,\varepsilon}_{\ell+1}<\xi\}, as well as the inclusion {σ~kK,ε=ξ}{σk+1K,ε=ξ}\{\tilde{\sigma}^{K,\varepsilon}_{k}=\xi\}\subset\{\sigma^{K,\varepsilon}_{k+1}=\xi\}. One easily concludes.

Step 3.2. For all K[[1,N]]K\subset[\![1,N]\!] such that |K|k0|K|\geq k_{0}, quasi-everywhere, uU{\mathbb{P}}^{U}_{u}-a.s., there is no t[0,ξ)t\in[0,\xi) such that RK(Ut)=0R_{K}(U_{t})=0. Indeed, on the contrary event, there is t[0,ξ)t\in[0,\xi) such that UtEk0U_{t}\notin E_{k_{0}}, whence Ut𝒰U_{t}\notin{\mathcal{U}}, which contradicts the fact that t[0,ξ)t\in[0,\xi).

Step 3.3. We show by decreasing induction that

𝒫(n): quasi-everywhere, uU-a.s. on the event {ξ=}bn=min{|K|=n}inft0RK(Ut)>0{\mathcal{P}}(n):\hbox{ \color[rgb]{0,0,0}quasi-everywhere, \color[rgb]{0,0,0}${\mathbb{P}}^{U}_{u}$-a.s. on the event $\{\xi=\infty\}$, $b_{n}=\min_{\{|K|=n\}}\inf_{t\geq 0}R_{K}(U_{t})>0$}

holds true for every n[[k0,N]]n\in[\![k_{0},N]\!].

The result is clear when n=Nn=N, because for all t[0,ξ)t\in[0,\xi), R[[1,N]](Ut)=1R_{[\![1,N]\!]}(U_{t})=1.

We next assume 𝒫(n){\mathcal{P}}(n) for some n[[k0+1,N]]n\in[\![k_{0}+1,N]\!] and we show that 𝒫(n1){\mathcal{P}}(n-1) is true. We fix K[[1,N]]K\subset[\![1,N]\!] with cardinal |K|=n1|K|=n-1 and we apply Step 3.1 with KK and with some ε(0,bn/(4aK))\varepsilon\in(0,b_{n}/(4a_{K})) (bnb_{n} is random but we may apply Step 3.1 simultaneously for all ε+(0,1]\varepsilon\in{\mathbb{Q}}_{+}^{*}\cap(0,1]) and Step 3.2, we find that on the event {ξ=}\{\xi=\infty\}, there either exists k1k\geq 1 such that σk+1K,ε=\sigma^{K,\varepsilon}_{k+1}=\infty or k1k\geq 1 such that inft[σkK,ε,σ~kK,ε)miniKRK{i}(Ut)2aKε\color[rgb]{0,0,0}\inf_{t\in[\sigma^{K,\varepsilon}_{k},\tilde{\sigma}^{K,\varepsilon}_{k})}\color[rgb]{0,0,0}\min_{i\notin K}R_{K\cup\{i\}}(U_{t})\leq 2a_{K}\varepsilon. This second choice is not possible, since by induction assumption, RK{i}(Ut)bnR_{K\cup\{i\}}(U_{t})\geq b_{n} for all t>0t>0 and all iKi\notin K. Hence there is k1k\geq 1 such that σk+1K,ε=\sigma^{K,\varepsilon}_{k+1}=\infty.

By definition of σk+1K,ε\sigma^{K,\varepsilon}_{k+1}, this implies that, still on the event where ξ=\xi=\infty, there exists t00t_{0}\geq 0 such that for all tt0t\geq t_{0}, either RK(Ut)ε/2R_{K}(U_{t})\geq\varepsilon/2 or miniKRK{i}(Ut)2aKε\min_{i\in K}R_{K\cup\{i\}}(U_{t})\leq 2a_{K}\varepsilon. Using again the induction assumption, we get that the second choice is never possible, so that actually, RK(Ut)ε/2R_{K}(U_{t})\geq\varepsilon/2 for all tt0t\geq t_{0}. Since (RK(Ut))t0(R_{K}(U_{t}))_{t\geq 0} is continuous and positive on [0,t0][0,t_{0}] according to Step 3.2, this completes the step.

Step 3.4. We conclude from Step 3.3 that quasi-everywhere, uU{\mathbb{P}}^{U}_{u}-a.s. on the event {ξ=}\{\xi=\infty\}, Ut𝒦U_{t}\in{\mathcal{K}} for all t0t\geq 0, where

𝒦={u𝒰:for all n[[k0,N]], all K[[1,N]] with |K|=nRK(u)bn}.{\mathcal{K}}=\{u\in{\mathcal{U}}:\hbox{for all $n\in[\![k_{0},N]\!]$, all $K\subset[\![1,N]\!]$ with $|K|=n$, $R_{K}(u)\geq b_{n}$}\}.

This (random) set is compact in 𝒰{\mathcal{U}}, so that Lemma 4.4-(i) tells us, both in the case where (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is recurrent and in the case where (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is transient, that this happens with probability 0. Hence quasi-everywhere, uU(ξ=)=0{\mathbb{P}}^{U}_{u}(\xi=\infty)=0 as desired. ∎

9.5. Binary collisions

We finally give the

Proof of Proposition 9.1-(iii).

We assume that N4N\geq 4, that 0<dθ,N(N)<2dθ,N(N1)0<d_{\theta,N}(N)<2\leq d_{\theta,N}(N-1) and observe that θ<2\theta<2 and k0>Nk_{0}>N, so that 𝒳=(2)N{\mathcal{X}}=({\mathbb{R}}^{2})^{N} and 𝒰=𝕊{\mathcal{U}}={\mathbb{S}}. The QKS(θ,N)QKS(\theta,N)-process 𝕏{\mathbb{X}} is non-exploding by Proposition 8.1-(i), and the QSKS(θ,N)QSKS(\theta,N)-process 𝕌{\mathbb{U}} is irreducible recurrent by Lemma 4.4-(ii). In particular, ζ=ξ=\zeta=\xi=\infty a.s. We divide the proof in 4 steps. First, we prove that 𝕏{\mathbb{X}} may have some binary collisions with positive probability. Then we check that this implies that 𝕌{\mathbb{U}} also may have some binary collisions with positive probability. Since 𝕌{\mathbb{U}} is recurrent, it will then necessarily be a.s. subjected to (infinitely many) binary collisions. Finally, we conclude using (2).

Step 1. We set 𝐊=({1,2},{3},,{N}){\mathbf{K}}=(\{1,2\},\{3\},\dots,\{N\}) and

𝒦={xB(0,C):x1x2<1 and mini[[1,N]],j[[3,N]],ijxixj>10},{\mathcal{K}}=\Big{\{}x\in B(0,C):\|x^{1}-x^{2}\|<1\;\mbox{ and }\;\min_{i\in[\![1,N]\!],j\in[\![3,N]\!],i\neq j}\|x^{i}-x^{j}\|>10\Big{\}},

with CC large enough so that μ(𝒦)>0\mu({\mathcal{K}})>0. We show in this step that xX(A)>0{\mathbb{P}}^{X}_{x}(A)>0 quasi-everywhere in 𝒦{\mathcal{K}}, where

A={Xt1=Xt2 for some t[0,1] and mint[0,1]R[[1,N]](Xt)>0}.A=\Big{\{}X^{1}_{t}=X^{2}_{t}\mbox{ for some }t\in[0,1]\;\mbox{ and }\;\min_{t\in[0,1]}R_{[\![1,N]\!]}(X_{t})>0\Big{\}}.

To this end, we fix x𝒦x\in{\mathcal{K}} and introduce the set

O={y(2)2:R{1,2}(y)<2,y1+y22x1+x22<1},\displaystyle O=\Big{\{}y\in({\mathbb{R}}^{2})^{2}:R_{\{1,2\}}(y)<2,\;\;\Big{\|}\frac{y^{1}+y^{2}}{2}-\frac{x^{1}+x^{2}}{2}\Big{\|}<1\Big{\}},

and Bi={y2:yxi2<1}B_{i}=\{y\in{\mathbb{R}}^{2}:||y-x^{i}||^{2}<1\} for i[[3,N]]i\in[\![3,N]\!]. Clearly, there is some ε(0,1]\varepsilon\in(0,1] such that

L={y(2)N:(y1,y2)O and yiBi for all i[[3,N]]}G𝐊,ε,L=\Big{\{}y\in({\mathbb{R}}^{2})^{N}:(y^{1},y^{2})\in O\;\hbox{ and }\;y^{i}\in B_{i}\hbox{ for all $i\in[\![3,N]\!]$}\Big{\}}\subset G_{{\mathbf{K}},\varepsilon},

where as usual G𝐊,ε={yB(0,1/ε):i[[1,N]],j[[3,N]]{i},yiyj2>ε}G_{{\mathbf{K}},\varepsilon}=\{y\in B(0,1/\varepsilon):\forall\;i\in[\![1,N]\!],\;\forall\;j\in[\![3,N]\!]\setminus\{i\},\;||y^{i}-y^{j}||^{2}>\varepsilon\}, recall that 𝒳=(2)N{\mathcal{X}}=({\mathbb{R}}^{2})^{N} because k0>Nk_{0}>N.

Since G𝐊,εG_{{\mathbf{K}},\varepsilon} is obviously included in {y(2)N:R[[1,N]](y)>0}\{y\in({\mathbb{R}}^{2})^{N}:R_{[\![1,N]\!]}(y)>0\}, we conclude that

xX(A)\displaystyle{\mathbb{P}}^{X}_{x}(A)\geq xX(Xt1=Xt2 for some t[0,1] and XtL for all t[0,1])\displaystyle{\mathbb{P}}^{X}_{x}\Big{(}X^{1}_{t}=X^{2}_{t}\mbox{ for some }t\in[0,1]\;\mbox{ and }\;X_{t}\in L\mbox{ for all }t\in[0,1]\Big{)}
\displaystyle\geq C1,ε,𝐊1x1,ε,𝐊(Xt1=Xt2 for some t[0,1] and XtL for all t[0,1])\displaystyle C_{1,\varepsilon,{\mathbf{K}}}^{-1}{\mathbb{Q}}^{1,\varepsilon,{\mathbf{K}}}_{x}\Big{(}X^{1}_{t}=X^{2}_{t}\mbox{ for some }t\in[0,1]\;\mbox{ and }\;X_{t}\in L\mbox{ for all }t\in[0,1]\Big{)}

by Proposition 7.1 with T=1T=1. We now set τ𝐊,ε=inf{t>0:XtG𝐊,ε}\tau_{{\mathbf{K}},\varepsilon}=\inf\{t>0:X_{t}\notin G_{{\mathbf{K}},\varepsilon}\}. Proposition 7.1 tells us that, quasi-everywhere in 𝒦G𝐊,ε{\mathcal{K}}\subset G_{{\mathbf{K}},\varepsilon}, the law of (Xt)t[0,τ𝐊,ε](X_{t})_{t\in[0,\tau_{{\mathbf{K}},\varepsilon}]} under x1,ε,𝐊{\mathbb{Q}}^{1,\varepsilon,{\mathbf{K}}}_{x} equals the law of Yt=(Yt1,,YtN)t[0,τ~𝐊,ε]Y_{t}=(Y^{1}_{t},\dots,Y^{N}_{t})_{t\in[0,\tilde{\tau}_{{\mathbf{K}},\varepsilon}]} where (Yt1,Yt2)t0(Y^{1}_{t},Y^{2}_{t})_{t\geq 0} is a QKS(2θ/N,2)QKS(2\theta/N,2)-process issued from (x1,x2)(x^{1},x^{2}), where for all i[[3,N]]i\in[\![3,N]\!], (Yti)t0(Y^{i}_{t})_{t\geq 0} is a QKS(θ/N,1)QKS(\theta/N,1)-process, i.e. a 22-dimensional Brownian motion, issued from xix^{i}, and where all these processes are independent. We have set τ~𝐊,ε=inf{t>0:YtG𝐊,ε}\tilde{\tau}_{{\mathbf{K}},\varepsilon}=\inf\{t>0:Y_{t}\notin G_{{\mathbf{K}},\varepsilon}\}. This implies, together with the fact that {XtL for all t[0,1]}{τ𝐊,ε>1}\{X_{t}\in L\mbox{ for all }t\in[0,1]\}\subset\{\tau_{{\mathbf{K}},\varepsilon}>1\}, that

xX(A)C1,ε,𝐊1pi=3Nqi{\mathbb{P}}^{X}_{x}(A)\geq C_{1,\varepsilon,{\mathbf{K}}}^{-1}\;p\prod_{i=3}^{N}q_{i}

quasi-everywhere in 𝒦{\mathcal{K}}, where

p=(mins[0,1]R{1,2}((Ys1,Ys2))=0 and (Yt1,Yt2)O for all t[0,1]),\displaystyle p={\mathbb{P}}\Big{(}\min_{s\in[0,1]}R_{\{1,2\}}((Y^{1}_{s},Y^{2}_{s}))=0\;\mbox{ and }\;(Y^{1}_{t},Y^{2}_{t})\in O\mbox{ for all }t\in[0,1]\Big{)},

and where qi=(YtiBiq_{i}={\mathbb{P}}(Y^{i}_{t}\in B_{i} for all t[0,1])t\in[0,1]). Of course, qi>0q_{i}>0 for all i[[3,N]]i\in[\![3,N]\!], since (Yti)t0(Y^{i}_{t})_{t\geq 0} is a Brownian motion issued from xix^{i}. Moreover, we know from Lemma 5.2 that (Mt=(Yt1+Yt2)/2)t0(M_{t}=(Y^{1}_{t}+Y^{2}_{t})/2)_{t\geq 0} is a 22-dimensional Brownian motion with diffusion coefficient 21/22^{-1/2} issued from m=(x1+x2)/2m=(x^{1}+x^{2})/2, that (Rt=R{1,2}((Yt1,Yt2)))t0(R_{t}=R_{\{1,2\}}((Y^{1}_{t},Y^{2}_{t})))_{t\geq 0} is a squared Bessel process of dimension d2θ/N,2(2)=dθ,N(2)d_{2\theta/N,2}(2)=d_{\theta,N}(2) issued from r=x1x22/2(0,1/2)r=||x^{1}-x^{2}||^{2}/2\in(0,1/2), and that these processes are independent. Hence, recalling the definition of OO,

p=(mins[0,1]Rs=0 and maxs[0,1]Rs<2)(maxs[0,1]Mtm<1).p={\mathbb{P}}\Big{(}\min_{s\in[0,1]}R_{s}=0\;\mbox{ and }\max_{s\in[0,1]}R_{s}<2\Big{)}{\mathbb{P}}\Big{(}\max_{s\in[0,1]}||M_{t}-m||<1\Big{)}.

This last quantity is clearly positive, because a squared Bessel process with dimension dθ,N(2)(0,2)d_{\theta,N}(2)\in(0,2), see Lemma 1.1, does hit zero, see Revuz-Yor [21, Chapter XI].

Step 2. We now deduce from Step 1 that the set F={u𝒰:u1=u2}F=\{u\in{\mathcal{U}}:u^{1}=u^{2}\} is not exceptional for 𝕌{\mathbb{U}}. Indeed, if it was exceptional, we would have uU(t0:UtF)=0{\mathbb{P}}^{U}_{u}(\exists\;t\geq 0:U_{t}\in F)=0 quasi-everywhere. By (2) and Remark 9.3, this would imply that quasi-everywhere, xX(t[0,τ):XtG)=0{\mathbb{P}}^{X}_{x}(\exists\;t\in[0,\tau):X_{t}\in G)=0, where G={x𝒳:x1=x2}G=\{x\in{\mathcal{X}}:x^{1}=x^{2}\} and τ=inf{t>0:R[[1,N]](Xt)=0}\tau=\inf\{t>0:R_{[\![1,N]\!]}(X_{t})=0\}. But on the event AA defined in Step 1, there is t[0,1]t\in[0,1] such that XtGX_{t}\in G and it holds that τ>1\tau>1. As a conclusion, xX(t[0,τ):XtG)>0{\mathbb{P}}^{X}_{x}(\exists\;t\in[0,\tau):X_{t}\in G)>0 quasi-everywhere in 𝒦{\mathcal{K}}, whence a contradiction, since μ(𝒦)>0\mu({\mathcal{K}})>0.

Step 3. Since (U,U)({\mathcal{E}}^{U},{\mathcal{F}}^{U}) is irreducible-recurrent and since FF is not exceptional, we know from Fukushima-Oshima-Takeda [11, Theorem 4.7.1-(iii) page 202] that quasi-everywhere,

uU(r>0,tr:UtF)=1.{\mathbb{P}}^{U}_{u}(\forall\;r>0,\exists\;t\geq r\;:\;U_{t}\in F)=1.

Step 4. Using again (2) and Remark 9.3 and recalling that ξ=\xi=\infty and that ρ\rho is an increasing bijection from [0,)[0,\infty) to [0,τ)[0,\tau), we conclude that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., XtX_{t} visits FF (an infinite number of times) during [0,τ)[0,\tau). Of course, the same arguments apply when replacing {1,2}\{1,2\} by any subset of [[1,N]][\![1,N]\!] with cardinal 22, and the proof is complete. ∎

10. Quasi-everywhere conclusion

Here we prove that the conclusions of Theorem 1.5 hold quasi-everywhere.

Partial proof of Theorem 1.5.

We assume that θ2\theta\geq 2 and N>3θN>3\theta, so that k0=2N/θ[[7,N]]k_{0}=\lceil 2N/\theta\rceil\in[\![7,N]\!], and consider a 𝒳{\mathcal{X}}_{\triangle}-valued QKS(θ,N)QKS(\theta,N)-process 𝕏{\mathbb{X}} with life-time ζ\zeta as in Proposition 4.1, where 𝒳=Ek0{\mathcal{X}}=E_{k_{0}}. Preliminaries. For K[[1,N]]K\subset[\![1,N]\!] and ε(0,1]\varepsilon\in(0,1], we write τK,ε=inf{t>0:XtGK,ε}[0,ζ]\tau_{K,\varepsilon}=\inf\{t>0:X_{t}\notin G_{K,\varepsilon}\}\in[0,\zeta] and GK,ε={x𝒳:miniK,jKxixj2>ε}B(0,1/ε)G_{K,\varepsilon}=\{x\in{\mathcal{X}}:\min_{i\in K,j\notin K}||x^{i}-x^{j}||^{2}>\varepsilon\}\cap B(0,1/\varepsilon) instead of τ𝐊,ε\tau_{{\mathbf{K}},\varepsilon} and G𝐊,εG_{{\mathbf{K}},\varepsilon} with 𝐊=(K,Kc){\mathbf{K}}=(K,K^{c}) as in Proposition 7.1. We also write xT,ε,K{\mathbb{Q}}^{T,\varepsilon,K}_{x} instead of T,ε,𝐊{\mathbb{Q}}^{T,\varepsilon,{\mathbf{K}}} and recall that it is equivalent to xX{\mathbb{P}}^{X}_{x} on TX=σ(Xs:s[0,T]){\mathcal{M}}^{X}_{T}=\sigma(X_{s}:s\in[0,T]).

Setting XtK=(Xti)iKX^{K}_{t}=(X^{i}_{t})_{i\in K} and XtKc=(Xti)iKcX^{K^{c}}_{t}=(X^{i}_{t})_{i\in K^{c}}, we know that quasi-everywhere in GK,εG_{K,\varepsilon}, the law of (XtK,XtKc)t[0,τK,εT](X^{K}_{t},X^{K^{c}}_{t})_{t\in[0,\tau_{K,\varepsilon}\land T]} under xT,ε,K{\mathbb{Q}}^{T,\varepsilon,K}_{x} is the same as the law of (Yt,Zt)t[0,τ~K,εT](Y_{t},Z_{t})_{t\in[0,\tilde{\tau}_{K,\varepsilon}\land T]}, where (Yt)t0(Y_{t})_{t\geq 0} is a QKS(|K|θ/N,|K|)QKS(|K|\theta/N,|K|)-process issued from x|Kx|_{K} and (Zt)t0(Z_{t})_{t\geq 0} is a QKS(|Kc|θ/N,|Kc|)QKS(|K^{c}|\theta/N,|K^{c}|)-process issued from x|Kcx|_{K^{c}}, these two processes being independent, and where τ~K,ε=inf{t>0:(Yt,Zt)GK,ε}\tilde{\tau}_{K,\varepsilon}=\inf\{t>0:(Y_{t},Z_{t})\notin G_{K,\varepsilon}\}. We denote by ζY\zeta^{Y} and ζZ\zeta^{Z} the life-times of (Yt)t0(Y_{t})_{t\geq 0} and (Zt)t0(Z_{t})_{t\geq 0}. The life-time of (Yt,Zt)t0(Y_{t},Z_{t})_{t\geq 0} is given by ζ=ζYζZ\zeta^{\prime}=\zeta^{Y}\land\zeta^{Z} and it holds that τ~K,ε[0,ζ]\tilde{\tau}_{K,\varepsilon}\in[0,\zeta^{\prime}].

No isolated points. Here we prove that for all K[[1,N]]K\subset[\![1,N]\!] with dθ,N(|K|)(0,2)d_{\theta,N}(|K|)\in(0,2), quasi-everywhere, we have xX(AK)=0{\mathbb{P}}^{X}_{x}(A_{K})=0, where AK={𝒵KA_{K}=\{{\mathcal{Z}}_{K} has an isolated point}\} and

𝒵K={t(0,ζ): there is a K-collision in the configuration Xt}.{\mathcal{Z}}_{K}=\{t\in(0,\zeta):\hbox{ there is a $K$-collision in the configuration }X_{t}\}.

On AKA_{K}, we can find u,v+u,v\in{\mathbb{Q}}_{+} such that u<v<ζu<v<\zeta and such that there is a unique t(u,v)t\in(u,v) with RK(Xt)=0R_{K}(X_{t})=0 and miniKRK{i}(Xt)>0\min_{i\notin K}R_{K\cup\{i\}}(X_{t})>0. By continuity, we deduce that on AKA_{K}, there exist r,s+r,s\in{\mathbb{Q}}_{+} and ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1] such that r<s<ζr<s<\zeta, XtGK,εX_{t}\in G_{K,\varepsilon} for all t[r,s]t\in[r,s] and such that {t(r,s):RK(Xt)=0}\{t\in(r,s):R_{K}(X_{t})=0\} has an isolated point. It thus suffices that for all r<sr<s and all ε(0,1]\varepsilon\in(0,1], that we all fix from now on, quasi-everywhere, xX(AK,r,s,ε)=0{\mathbb{P}}^{X}_{x}(A_{K,r,s,\varepsilon})=0, where

AK,r,s,ε={XtGK,ε for all t(r,s) and {t(r,s):RK(Xt)=0} has an isolated point}.\displaystyle A_{K,r,s,\varepsilon}=\Big{\{}\hbox{$X_{t}\in G_{K,\varepsilon}$ for all $t\in(r,s)$ and $\{t\in(r,s):R_{K}(X_{t})=0\}$ has an isolated point}\Big{\}}.

By the Markov property, it suffices that xX(AK,0,s,ε)=0{\mathbb{P}}^{X}_{x}(A_{K,0,s,\varepsilon})=0 quasi-everywhere in GK,εG_{K,\varepsilon} and, by equivalence, that xs,ε,K(AK,0,s,ε)=0{\mathbb{Q}}^{s,\varepsilon,K}_{x}(A_{K,0,s,\varepsilon})=0 quasi-everywhere in GK,εG_{K,\varepsilon}. We write, recalling the preliminaries,

xs,ε,K(AK,0,s,ε)=\displaystyle{\mathbb{Q}}^{s,\varepsilon,K}_{x}(A_{K,0,s,\varepsilon})= xs,ε,K(τK,εs and {t(0,s):RK(Xt)=0} has an isolated point)\displaystyle{\mathbb{Q}}^{s,\varepsilon,K}_{x}\Big{(}\tau_{K,\varepsilon}\geq s\hbox{ and $\{t\in(0,s):R_{K}(X_{t})=0\}$ has an isolated point}\Big{)}
=\displaystyle= (τ~K,εs and {t(0,s):RK(Yt)=0} has an isolated point)\displaystyle{\mathbb{P}}\Big{(}\tilde{\tau}_{K,\varepsilon}\geq s\hbox{ and $\{t\in(0,s):R_{K}(Y_{t})=0\}$ has an isolated point}\Big{)}
\displaystyle\leq ( {t(0,s):RK(Yt)=0} has an isolated point).\displaystyle{\mathbb{P}}\Big{(}\hbox{ $\{t\in(0,s):R_{K}(Y_{t})=0\}$ has an isolated point}\Big{)}.

But (Yt)t0(Y_{t})_{t\geq 0} is a QKS(|K|θ/N,|K|)QKS(|K|\theta/N,|K|)-process, so that we know from Lemma 5.2 that (RK(Yt))t0(R_{K}(Y_{t}))_{t\geq 0} is a squared Bessel process with dimension d|K|θ/N,|K|(|K|)=dθ,N(|K|)(0,2)d_{|K|\theta/N,|K|}(|K|)=d_{\theta,N}(|K|)\in(0,2). Such a process has no isolated zero, see Revuz-Yor [21, Chapter XI].

Point (i). We have already seen in Proposition 8.1-(ii) that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., ζ<\zeta<\infty and Xζ=limtζXtX_{\zeta-}=\lim_{t\to\zeta-}X_{t} exists in (2)N({\mathbb{R}}^{2})^{N} and does not belong to Ek0E_{k_{0}}.

Point (ii). We want to show that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., there is K0[[1,N]]K_{0}\subset[\![1,N]\!] with |K0|=k0|K_{0}|=k_{0} such that there is a K0K_{0}-collision and no KK-collision with |K|>k0|K|>k_{0} in the configuration XζX_{\zeta-}. We already know that XζEk0X_{\zeta-}\notin E_{k_{0}}, so that there is K[[1,N]]K\subset[\![1,N]\!] with |K|k0|K|\geq k_{0} such that there is a KK-collision in the configuration XζX_{\zeta-}. Hence the goal is to verify that quasi-everywhere, for all K[[1,N]]K\subset[\![1,N]\!] with |K|>k0|K|>k_{0}, xX(BK)=0{\mathbb{P}}^{X}_{x}(B_{K})=0, where

BK={There is a K-collision in the configuration Xζ}.B_{K}=\{\hbox{There is a $K$-collision in the configuration }X_{\zeta-}\}.

On BKB_{K}, there is ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1] such that XζGK,2εX_{\zeta-}\in G_{K,2\varepsilon}. By continuity, there also exists, still on BKB_{K}, some r+[0,ζ)r\in{\mathbb{Q}}_{+}\cap[0,\zeta) such that XtGK,εX_{t}\in G_{K,\varepsilon} for all t[r,ζ)t\in[r,\zeta). Hence we only have to prove that for all ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1], all t+t\in{\mathbb{Q}}_{+}, all T+T\in{\mathbb{Q}}_{+} such that T>rT>r, quasi-everywhere, xX(BK,r,T,ε)=0{\mathbb{P}}^{X}_{x}(B_{K,r,T,\varepsilon})=0, where

BK,r,T,ε={ζ(r,T],XtGK,ε for all t[r,ζ) and RK(Xζ)=0}.B_{K,r,T,\varepsilon}=\{\zeta\in(r,T],\;X_{t}\in G_{K,\varepsilon}\hbox{ for all $t\in[r,\zeta)$ and }R_{K}(X_{\zeta-})=0\}.

By the Markov property, it suffices that xX(BK,0,T,ε)=0{\mathbb{P}}^{X}_{x}(B_{K,0,T,\varepsilon})=0 quasi-everywhere in GK,εG_{K,\varepsilon}, for all ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1] and all T+T\in{\mathbb{Q}}_{+}^{*}. We now fix ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1] and T+T\in{\mathbb{Q}}_{+}^{*}. By equivalence, it suffices to prove that xT,ε,K(BK,0,T,ε)=0{\mathbb{Q}}^{T,\varepsilon,K}_{x}(B_{K,0,T,\varepsilon})=0. Using the notation introduced in the preliminaries, we write

xT,ε,K(BK,0,T,ε)=\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}(B_{K,0,T,\varepsilon})= xT,ε,K(ζT,τK,ε=ζ and RK(Xζ)=0)\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}\Big{(}\zeta\leq T,\;\tau_{K,\varepsilon}=\zeta\hbox{ and }R_{K}(X_{\zeta-})=0\Big{)}
=\displaystyle= (ζT,τ~K,ε=ζ and RK(Yζ)=0)\displaystyle{\mathbb{P}}\Big{(}\zeta^{\prime}\leq T,\;\tilde{\tau}_{K,\varepsilon}=\zeta^{\prime}\hbox{ and }R_{K}(Y_{\zeta^{\prime}-})=0\Big{)}
\displaystyle\leq (inft[0,ζY)RK(Yt)=0).\displaystyle{\mathbb{P}}\Big{(}\inf_{t\in[0,\zeta^{Y})}R_{K}(Y_{t})=0\Big{)}.

But (Yt)t0(Y_{t})_{t\geq 0} is a QKS(|K|θ/N,|K|)QKS(|K|\theta/N,|K|)-process with |K|>k07|K|>k_{0}\geq 7 and with d|K|θ/N,|K|(|K|1)=dθ,N(|K|1)0d_{|K|\theta/N,|K|}(|K|-1)=d_{\theta,N}(|K|-1)\leq 0 by Lemma 1.1 because |K|1k0|K|-1\geq k_{0}. We also have d|K|θ/N,|K|(|K|)=dθ,N(|K|)0d_{|K|\theta/N,|K|}(|K|)=d_{\theta,N}(|K|)\leq 0. Hence Proposition 9.1-(i) tells us that (inft[0,ζY)RK(Yt)=0)=0{\mathbb{P}}(\inf_{t\in[0,\zeta^{Y})}R_{K}(Y_{t})=0)=0.

Point (iii). We recall that k1=k01k_{1}=k_{0}-1 and we fix LK[[1,N]]L\subset K\subset[\![1,N]\!] with |K|=k0|K|=k_{0} and |L|=k1|L|=k_{1}. We want to prove that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., if RK(Xζ)=0R_{K}(X_{\zeta-})=0, then for all t[0,ζ)t\in[0,\zeta), the set 𝒵L(t,ζ){\mathcal{Z}}_{L}\cap(t,\zeta) is infinite and has no isolated point. But since dθ,N(k1)(0,2)d_{\theta,N}(k_{1})\in(0,2), see Lemma 1.1, we already know that 𝒵L{\mathcal{Z}}_{L} has no isolated point. It thus suffices to check that quasi-everywhere, for all r+r\in{\mathbb{Q}}_{+}, we have xX(CK,L,r)=0{\mathbb{P}}^{X}_{x}(C_{K,L,r})=0, where

CK,L,r={ζ>r,RK(Xζ)=0, and RL(Xt)>0 for all t(r,ζ)}.C_{K,L,r}=\{\zeta>r,\;R_{K}(X_{\zeta-})=0,\hbox{ and }R_{L}(X_{t})>0\hbox{ for all }t\in(r,\zeta)\}.

We used that since |L|=k1=k01|L|=k_{1}=k_{0}-1, for all x𝒳=Ek0x\in{\mathcal{X}}=E_{k_{0}}, there is a LL collision in the configuration xx if and only if RL(x)=0R_{L}(x)=0. On CK,L,rC_{K,L,r}, thanks to point (ii) , there are ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1], T+T\in{\mathbb{Q}}_{+} and s+[r,ζ)s\in{\mathbb{Q}}_{+}^{*}\cap[r,\zeta) such that ζ(s,T]\zeta\in(s,T] and XtGK,εX_{t}\in G_{K,\varepsilon} for all t[s,ζ)t\in[s,\zeta). Thus it suffices to prove that for all s<Ts<T and all ε(0,1]\varepsilon\in(0,1], that we now fix, quasi-everywhere, xX(CK,L,s,T,ε)=0{\mathbb{P}}^{X}_{x}(C_{K,L,s,T,\varepsilon})=0, where

CK,L,s,T,ε={ζ(s,T],RK(Xζ)=0,XtGK,ε and RL(Xt)>0 for all t[s,ζ)}.C_{K,L,s,T,\varepsilon}=\{\zeta\in(s,T],\;R_{K}(X_{\zeta-})=0,\;X_{t}\in G_{K,\varepsilon}\hbox{ and }R_{L}(X_{t})>0\hbox{ for all }t\in[s,\zeta)\}.

By the Markov property, it suffices that xX(CK,L,0,T,ε)=0{\mathbb{P}}^{X}_{x}(C_{K,L,0,T,\varepsilon})=0 quasi-everywhere in GK,εG_{K,\varepsilon} and, by equivalence, that xT,ε,K(CK,L,0,T,ε)=0{\mathbb{Q}}^{T,\varepsilon,K}_{x}(C_{K,L,0,T,\varepsilon})=0. Recalling the preliminaries, we write

xT,ε,K(CK,L,0,T,ε)=\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}(C_{K,L,0,T,\varepsilon})= xT,ε,K(ζT,RK(Xζ)=0,τK,ε=ζ and RL(Xt)>0 for all t[0,ζ))\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}\Big{(}\zeta\leq T,\;R_{K}(X_{\zeta-})=0,\;\tau_{K,\varepsilon}=\zeta\hbox{ and }R_{L}(X_{t})>0\hbox{ for all }t\in[0,\zeta)\Big{)}
=\displaystyle= (ζT,RK(Yζ)=0,τ~K,ε=ζ and RL(Yt)>0 for all t[0,ζ)).\displaystyle{\mathbb{P}}\Big{(}\zeta^{\prime}\leq T,\;R_{K}(Y_{\zeta^{\prime}-})=0,\;\tilde{\tau}_{K,\varepsilon}=\zeta^{\prime}\hbox{ and }R_{L}(Y_{t})>0\hbox{ for all }t\in[0,\zeta^{\prime})\Big{)}.

Setting σK=inf{t>0:RK(Yt)=0}\sigma_{K}=\inf\{t>0:R_{K}(Y_{t})=0\}, we observe that σK=ζY\sigma_{K}=\zeta^{Y}. Indeed, |K|=k0|K|=k_{0} and (Yt)t0(Y_{t})_{t\geq 0} is a QKS(|K|θ/N,|K|)QKS(|K|\theta/N,|K|)-process, of which the state space is given by 𝒴=𝒴{}{\mathcal{Y}}_{\triangle}={\mathcal{Y}}\cup\{\triangle\}, where 𝒴={y(2)|K|:RM(y)>0{\mathcal{Y}}=\{y\in({\mathbb{R}}^{2})^{|K|}:R_{M}(y)>0 for all M[[1,N]]M\subset[\![1,N]\!] such that |M|k0}|M|\geq k_{0}\}, because 2|K|/(|K|θ/N)=2N/θ=k0\lceil 2|K|/(|K|\theta/N)\rceil=\lceil 2N/\theta\rceil=k_{0}. Hence {RK(Yζ)=0}{ζ=σK}\{R_{K}(Y_{\zeta^{\prime}-})=0\}\subset\{\zeta^{\prime}=\sigma_{K}\}, so that

xT,ε,K(CK,L,0,T,ε)(RL(Yt)>0 for all t[0,σK)).\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}(C_{K,L,0,T,\varepsilon})\leq{\mathbb{P}}(R_{L}(Y_{t})>0\hbox{ for all }t\in[0,\sigma_{K})).

This last quantity equals zero by Proposition 9.1-(ii), since d|K|θ/N,|K|(|K|1)=dθ,N(|K|1)=dθ,N(k01)(0,2)d_{|K|\theta/N,|K|}(|K|-1)=d_{\theta,N}(|K|-1)=d_{\theta,N}(k_{0}-1)\in(0,2) by Lemma 1.1 and since |L|=k1=|K|1|L|=k_{1}=|K|-1 and since d|K|θ/N,|K|(|K|)=dθ,N(|K|)=dθ,N(k0)0<2d_{|K|\theta/N,|K|}(|K|)=d_{\theta,N}(|K|)=d_{\theta,N}(k_{0})\leq 0<2.

Point (iv). We assume that k2=k02k_{2}=k_{0}-2, i.e. that dθ,N(k02)(0,2)d_{\theta,N}(k_{0}-2)\in(0,2). We fix LK[[1,N]]L\subset K\subset[\![1,N]\!] with |K|=k1|K|=k_{1} and |L|=k2|L|=k_{2}. We want to prove that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., for all t[0,ζ)t\in[0,\zeta), if there is a KK-collision in the configuration XtX_{t}, then for all r[0,t)r\in[0,t), the set 𝒵L(r,t){\mathcal{Z}}_{L}\cap(r,t) is infinite and has no isolated point. We already know that 𝒵L{\mathcal{Z}}_{L} has no isolated point. It thus suffices to check that quasi-everywhere, for all r+r\in{\mathbb{Q}}_{+}, we have xX(DK,L,r)=0{\mathbb{P}}^{X}_{x}(D_{K,L,r})=0, where

DK,L,r={ζ>r and there is t(r,ζ) such that there is a K-collision at time t\displaystyle D_{K,L,r}=\{\zeta>r\hbox{ and there is $t\in(r,\zeta)$ such that there is a $K$-collision at time $t$}
but no L-collision during (r,t)}.\displaystyle\hskip 256.0748pt\hbox{but no $L$-collision during $(r,t)$}\}.

We set σK,r=inf{t>r:\sigma_{K,r}=\inf\{t>r: there is a KK-collision in the configuration Xt}X_{t}\}. It holds that

DK,L,r={ζ>r,σK,r<ζ and there is no L-collision during u[r,σK,r)}.D_{K,L,r}=\{\zeta>r,\;\sigma_{K,r}<\zeta\hbox{ and there is no $L$-collision during }u\in[r,\sigma_{K,r})\}.

On DK,L,rD_{K,L,r}, there exists ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1] such that XσK,rGK,2εX_{\sigma_{K,r}}\in G_{K,2\varepsilon}, so that by continuity, there exists v+[r,σK,r)v\in{\mathbb{Q}}_{+}\cap[r,\sigma_{K,r}) such that XuGK,εX_{u}\in G_{K,\varepsilon} for all u[v,σK,r]u\in[v,\sigma_{K,r}]. Observe that σK,v=σK,r\sigma_{K,v}=\sigma_{K,r} and that for all t[v,σK,v)t\in[v,\sigma_{K,v}), there is a LL-collision at time tt if and only if RL(Xt)=0R_{L}(X_{t})=0, by definition of σK,v\sigma_{K,v} and since XtGK,εX_{t}\in G_{K,\varepsilon}. All in all, it suffices to prove that for all v+v\in{\mathbb{Q}}_{+}, all ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1], all T+T\in{\mathbb{Q}}_{+}^{*}, xX(DK,L,v,T,ε)=0{\mathbb{P}}^{X}_{x}(D_{K,L,v,T,\varepsilon})=0 quasi-everywhere, where

DK,L,v,T,ε={ζ(v,T],σK,v<ζ,XuGK,ε and RL(Xu)>0 for all u[v,σK,v)}.D_{K,L,v,T,\varepsilon}=\{\zeta\in(v,T],\;\sigma_{K,v}<\zeta,\;X_{u}\in G_{K,\varepsilon}\hbox{ and }R_{L}(X_{u})>0\hbox{ for all }u\in[v,\sigma_{K,v})\}.

By the Markov property, it suffices to prove that xX(DK,L,0,T,ε)=0{\mathbb{P}}^{X}_{x}(D_{K,L,0,T,\varepsilon})=0 quasi-everywhere in GK,εG_{K,\varepsilon} and, by equivalence, we may use xT,ε,K{\mathbb{Q}}^{T,\varepsilon,K}_{x} instead of xX{\mathbb{P}}^{X}_{x}. But recalling the preliminaries,

xT,ε,K(DK,L,0,T,ε)=\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}(D_{K,L,0,T,\varepsilon})= xT,ε,K(ζT,σK,0<ζ,τK,εσK,0 and RL(Xt)>0 for all t[0,σK,0))\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}\Big{(}\zeta\leq T,\;\sigma_{K,0}<\zeta,\;\tau_{K,\varepsilon}\geq\sigma_{K,0}\;\hbox{ and }\;R_{L}(X_{t})>0\hbox{ for all }t\in[0,\sigma_{K,0})\Big{)}
=\displaystyle= (ζT,σ~K,0<ζ,τ~K,εσ~K,0 and RL(Yt)>0 for all t[0,σ~K,0))\displaystyle{\mathbb{P}}\Big{(}\zeta^{\prime}\leq T,\;\tilde{\sigma}_{K,0}<\zeta^{\prime},\;\tilde{\tau}_{K,\varepsilon}\geq\tilde{\sigma}_{K,0}\;\hbox{ and }\;R_{L}(Y_{t})>0\hbox{ for all }t\in[0,\tilde{\sigma}_{K,0})\Big{)}
\displaystyle\leq (RL(Yt)>0 for all t[0,σ~K,0)),\displaystyle{\mathbb{P}}\Big{(}R_{L}(Y_{t})>0\hbox{ for all }t\in[0,\tilde{\sigma}_{K,0})\Big{)},

where we have set σ~K,0=inf{t>0:RK(Yt)=0}\tilde{\sigma}_{K,0}=\inf\{t>0:R_{K}(Y_{t})=0\}. Finally, (RL(Yt)>0 for all t[0,σ~K,0))=0{\mathbb{P}}(R_{L}(Y_{t})>0\hbox{ for all }t\in[0,\tilde{\sigma}_{K,0}))=0 by Proposition 9.1-(ii), because (Yt)t0(Y_{t})_{t\geq 0} is a QKS(|K|θ/N,|K|)QKS(|K|\theta/N,|K|)-process, because |L|=k2=|K|1|L|=k_{2}=|K|-1, because d|K|θ/N,|K|(|K|1)=dθ,N(|K|1)=dθ,N(k2)(0,2)d_{|K|\theta/N,|K|}(|K|-1)=d_{\theta,N}(|K|-1)=d_{\theta,N}(k_{2})\in(0,2) and because d|K|θ/N,|K|(|K|)=dθ,N(|K|)=dθ,N(k1)(0,2)d_{|K|\theta/N,|K|}(|K|)=d_{\theta,N}(|K|)=d_{\theta,N}(k_{1})\in(0,2).

Point (v). We fix K[[1,N]]K\subset[\![1,N]\!] with cardinal |K|[[3,k21]]|K|\in[\![3,k_{2}-1]\!], so that dθ,N(|K|)2d_{\theta,N}(|K|)\geq 2. We want to prove that quasi-everywhere, xX{\mathbb{P}}^{X}_{x}-a.s., for all t[0,ζ)t\in[0,\zeta), there is no KK-collision in the configuration XtX_{t}. We introduce σK=inf{t>0:\sigma_{K}=\inf\{t>0: there is a KK-collision in the configuration Xt}X_{t}\}, with the convention that inf=ζ\inf\emptyset=\zeta, and we have to verify that quasi-everywhere, xX(σK<ζ)=0{\mathbb{P}}^{X}_{x}(\sigma_{K}<\zeta)=0.

On the event {σK<ζ}\{\sigma_{K}<\zeta\}, there exist ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1] and r+[0,σK)r\in{\mathbb{Q}}_{+}^{*}\cap[0,\sigma_{K}) such that XtGK,εX_{t}\in G_{K,\varepsilon} for all t[r,σK]t\in[r,\sigma_{K}]. Hence it suffices to check that for all ε(0,1]\varepsilon\in{\mathbb{Q}}\cap(0,1], all r+r\in{\mathbb{Q}}_{+}^{*} and all T+(r,)T\in{\mathbb{Q}}_{+}^{*}\cap(r,\infty), which we now fix, quasi-everywhere, xX(FK,r,T,ε)=0{\mathbb{P}}^{X}_{x}(F_{K,r,T,\varepsilon})=0, where

FK,r,T,ε={σK(r,ζT) and XtGK,ε for all t[r,σK]}.F_{K,r,T,\varepsilon}=\{\sigma_{K}\in(r,\zeta\land T)\hbox{ and $X_{t}\in G_{K,\varepsilon}$ for all $t\in[r,\sigma_{K}]$}\}.

By the Markov property, it suffices that xX(FK,0,T,ε)=0{\mathbb{P}}^{X}_{x}(F_{K,0,T,\varepsilon})=0 quasi-everywhere in GK,εG_{K,\varepsilon} and, by equivalence, that xT,ε,K(FK,0,T,ε)=0{\mathbb{Q}}^{T,\varepsilon,K}_{x}(F_{K,0,T,\varepsilon})=0. Recalling the preliminaries, we write

xT,ε,K(FK,0,T,ε)=\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}(F_{K,0,T,\varepsilon})= xT,ε,K(σK(0,ζT) and τK,εσK)\displaystyle{\mathbb{Q}}^{T,\varepsilon,K}_{x}\Big{(}\sigma_{K}\in(0,\zeta\land T)\hbox{ and }\tau_{K,\varepsilon}\geq\sigma_{K}\Big{)}
=\displaystyle= (σ~K(0,ζT) and τ~K,εσ~K)\displaystyle{\mathbb{P}}\Big{(}\tilde{\sigma}_{K}\in(0,\zeta^{\prime}\land T)\hbox{ and }\tilde{\tau}_{K,\varepsilon}\geq\tilde{\sigma}_{K}\Big{)}
\displaystyle\leq (inft[0,T]RK(Yt)=0),\displaystyle{\mathbb{P}}\Big{(}\inf_{t\in[0,T]}R_{K}(Y_{t})=0\Big{)},

where we have set σ~K=inf{t>0:\tilde{\sigma}_{K}=\inf\{t>0: there is a KK-collision in the configuration (Yt,Zt)}(Y_{t},Z_{t})\}. Since (Yt)t0(Y_{t})_{t\geq 0} is a QKS(|K|θ/N,|K|)QKS(|K|\theta/N,|K|)-process, we know from Lemma 5.2 that (RK(Yt))t0(R_{K}(Y_{t}))_{t\geq 0} is a squared Bessel process with dimension d|K|θ/N,|K|(|K|)=dθ,N(|K|)2d_{|K|\theta/N,|K|}(|K|)=d_{\theta,N}(|K|)\geq 2. Such a process does a.s. never reach 0.

Point (vi). The proof is exactly the same as that of (iv), replacing everywhere k1k_{1} by k2k_{2} and k2k_{2} by 22, and using Proposition 9.1-(iii) instead of Proposition 9.1-(ii), which is licit because 0<dk2θ/N,k2(k2)<2dk2θ/N,k2(k21){0<d_{k_{2}\theta/N,k_{2}}(k_{2})<2\leq d_{k_{2}\theta/N,k_{2}}(k_{2}-1)}, since dk2θ/N,k2(k2)=dθ,N(k2)d_{k_{2}\theta/N,k_{2}}(k_{2})=d_{\theta,N}(k_{2}) and dk2θ/N,k2(k21)=dθ,N(k21)d_{k_{2}\theta/N,k_{2}}(k_{2}-1)=d_{\theta,N}(k_{2}-1) and by Lemma 1.1. ∎

11. Extension to all initial conditions in E2E_{2}

We first prove Proposition 1.2: we can build a KS(θ,N)KS(\theta,N)-process, i.e. a QKS(θ,N)QKS(\theta,N)-process such that xXXt1{\mathbb{P}}_{x}^{X}\circ X_{t}^{-1} is absolutely continuous for all xE2x\in E_{2} and all t>0t>0. We next conclude the proofs of Proposition 1.3 and of Theorem 1.5.

11.1. Construction of a KS(θ,N)KS(\theta,N)-process

We fix θ>0\theta>0 and N2N\geq 2 such that N>θN>\theta during the whole subsection. For each nn\in{\mathbb{N}}^{*}, we introduce ϕnC(+,+)\phi_{n}\in C^{\infty}({\mathbb{R}}_{+},{\mathbb{R}}_{+}^{*}) such that ϕn(r)=r\phi_{n}(r)=r for all r1/nr\geq 1/n and we set, for x(2)Nx\in({\mathbb{R}}^{2})^{N},

𝐦n(x)=1ijN[ϕn(xixj2)]θ/Nandμn(dx)=𝐦n(x)dx.{\mathbf{m}}_{n}(x)=\prod_{1\leq i\neq j\leq N}[\phi_{n}(\|x^{i}-x^{j}\|^{2})]^{-\theta/N}\qquad\hbox{and}\qquad\mu_{n}({\rm d}x)={\mathbf{m}}_{n}(x){\rm d}x.

We then consider the (2)N({\mathbb{R}}^{2})^{N}-valued S.D.E

(1) Xtn=x+Bt+0t𝐦n(Xsn)2𝐦n(Xsn)ds,\displaystyle X^{n}_{t}=x+B_{t}+\int_{0}^{t}\frac{\nabla{\mathbf{m}}_{n}(X^{n}_{s})}{2{\mathbf{m}}_{n}(X^{n}_{s})}{\rm d}s,

which is strongly well-posed, for every initial condition, since the drift coefficient is smooth and bounded. We denote by 𝕏n=(Ωn,n,(Xtn)t0,(xn)x(2)N){\mathbb{X}}^{n}=(\Omega^{n},{\mathcal{M}}^{n},(X^{n}_{t})_{t\geq 0},({\mathbb{P}}^{n}_{x})_{x\in({\mathbb{R}}^{2})^{N}}) the corresponding Markov process.

Lemma 11.1.

For all n1n\geq 1, 𝕏n{\mathbb{X}}^{n} is a μn\mu_{n}-symmetric (2)N({\mathbb{R}}^{2})^{N}-valued diffusion with regular Dirichlet space (n,n)({\mathcal{E}}^{n},{\mathcal{F}}^{n}) with core Cc((2)N)C_{c}^{\infty}(({\mathbb{R}}^{2})^{N}) such that for all φCc((2)N)\varphi\in C_{c}^{\infty}(({\mathbb{R}}^{2})^{N}),

n(φ,φ)=12(2)Nφ2dμn.{\mathcal{E}}^{n}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\|\nabla\varphi\|^{2}{\rm d}\mu_{n}.

Moreover xn(Xtn)1{\mathbb{P}}^{n}_{x}\circ(X_{t}^{n})^{-1} has a density with respect to the Lebesgue measure on (2)N({\mathbb{R}}^{2})^{N} for all t>0t>0 and all x(2)Nx\in({\mathbb{R}}^{2})^{N}.

Proof.

Classically, 𝕏n{\mathbb{X}}^{n} is a μn\mu_{n}-symmetric diffusion and its (strong) generator n{\mathcal{L}}^{n} satisfies that for all φCc((2)N)\varphi\in C^{\infty}_{c}(({\mathbb{R}}^{2})^{N}), all x(2)Nx\in({\mathbb{R}}^{2})^{N}, nφ(x)=12Δφ(x)+𝐦n(x)2𝐦n(x)φ(x){\mathcal{L}}^{n}\varphi(x)=\frac{1}{2}\Delta\varphi(x)+\frac{\nabla{\mathbf{m}}_{n}(x)}{2{\mathbf{m}}_{n}(x)}\cdot\nabla\varphi(x). Hence, see Subsection B.1, one easily shows that for (n,n)({\mathcal{E}}^{n},{\mathcal{F}}^{n}) the Dirichlet space of 𝕏n{\mathbb{X}}^{n}, we have Cc((2)N)nC_{c}^{\infty}(({\mathbb{R}}^{2})^{N})\subset{\mathcal{F}}^{n} and, for φCc((2)N)\varphi\in C_{c}^{\infty}(({\mathbb{R}}^{2})^{N}), n(φ,φ)=12(2)Nφ2dμn{\mathcal{E}}^{n}(\varphi,\varphi)=\frac{1}{2}\int_{({\mathbb{R}}^{2})^{N}}\|\nabla\varphi\|^{2}{\rm d}\mu_{n}. Since (n,n)({\mathcal{E}}^{n},{\mathcal{F}}^{n}) is closed, we deduce that

Cc((2)N)¯1nn,\overline{C_{c}^{\infty}(({\mathbb{R}}^{2})^{N})}^{{\mathcal{E}}^{n}_{1}}\subset{\mathcal{F}}^{n},

where 1n(,)=n(,)+L2((2)N,μn)2{\mathcal{E}}^{n}_{1}(\cdot,\cdot)={\mathcal{E}}^{n}(\cdot,\cdot)+\|\cdot\|^{2}_{L^{2}(({\mathbb{R}}^{2})^{N},\mu_{n})}. But thanks to [11, Lemma 3.3.5 page 136],

n{φL2((2)N,μn):φL2((2)N,μn)},{\mathcal{F}}^{n}\subset\{\varphi\in L^{2}(({\mathbb{R}}^{2})^{N},\mu_{n}):\nabla\varphi\in L^{2}(({\mathbb{R}}^{2})^{N},\mu_{n})\},

where \nabla is understood in the sense of distributions. Since finally

Cc((2)N)¯1n={φL2((2)N,μn):φL2((2)N,μn)},\overline{C_{c}^{\infty}(({\mathbb{R}}^{2})^{N})}^{{\mathcal{E}}^{n}_{1}}=\{\varphi\in L^{2}(({\mathbb{R}}^{2})^{N},\mu_{n}):\nabla\varphi\in L^{2}(({\mathbb{R}}^{2})^{N},\mu_{n})\},

𝕏n{\mathbb{X}}^{n} has the announced Dirichlet space. Finally, the absolute continuity of xn(Xtn)1{\mathbb{P}}^{n}_{x}\circ(X_{t}^{n})^{-1}, for t>0t>0 and x(2)Nx\in({\mathbb{R}}^{2})^{N}, immediately follows from the (standard) Girsanov theorem, since the drift coefficient is bounded. ∎

For all xE2x\in E_{2} we set dx=minijxixj2d_{x}=\min_{i\neq j}\|x^{i}-x^{j}\|^{2}. For n1n\geq 1, we introduce the open set

(2) E2n={x(2)N:dx>1n and x<n}.E_{2}^{n}=\Big{\{}x\in({\mathbb{R}}^{2})^{N}:d_{x}>\frac{1}{n}\;\hbox{ and }\;||x||<n\Big{\}}.

We also fix a QKS(θ,N)QKS(\theta,N)-process 𝕏=(ΩX,X,(Xt)t0,(xX)x𝒳){\mathbb{X}}=(\Omega^{X},{\mathcal{M}}^{X},(X_{t})_{t\geq 0},({\mathbb{P}}^{X}_{x})_{x\in{\mathcal{X}}_{\triangle}}) for the whole subsection.

Lemma 11.2.

There exists an exceptional set 𝒩0E2{\mathcal{N}}_{0}\subset E_{2} with respect to 𝕏{\mathbb{X}} such that for all n1n\geq 1, for all xE2n𝒩0x\in E_{2}^{n}\setminus{\mathcal{N}}_{0}, the law of (Xtτnn)t0(X^{n}_{t\land\tau_{n}})_{t\geq 0} under xn{\mathbb{P}}^{n}_{x} equals the law of (Xtσn)t0(X_{t\land\sigma_{n}})_{t\geq 0} under xX{\mathbb{P}}^{X}_{x}, where

τn=inf{t>0:XtnE2n}andσn=inf{t>0:XtE2n}.\tau_{n}=\inf\{t>0:X_{t}^{n}\notin E^{n}_{2}\}\qquad\hbox{and}\qquad\sigma_{n}=\inf\{t>0:X_{t}\notin E^{n}_{2}\}.
Proof.

We fix n1n\geq 1. Applying Lemma B.6 to 𝕏n{\mathbb{X}}^{n} and 𝕏{\mathbb{X}} with the open set E2nE^{n}_{2}, using that 𝐦n=𝐦{\mathbf{m}}_{n}={\mathbf{m}} on E2nE_{2}^{n} and Lemma 11.1, we find that the processes 𝕏n{\mathbb{X}}^{n} and 𝕏{\mathbb{X}} killed when leaving E2nE^{n}_{2} have the same Dirichlet space. By uniqueness, see [11, Theorem 4.2.8 page 167], there exists an exceptional set 𝒩n{\mathcal{N}}_{n} such that for all xE2n𝒩nx\in E^{n}_{2}\setminus{\mathcal{N}}_{n}, the law of (Xtn)t0(X^{n}_{t})_{t\geq 0} killed when leaving E2nE^{n}_{2} under xn{\mathbb{P}}^{n}_{x} equals the law of (Xt)t0(X_{t})_{t\geq 0} killed when leaving E2nE^{n}_{2} under xX{\mathbb{P}}^{X}_{x}. We conclude setting 𝒩0=n1𝒩n{\mathcal{N}}_{0}=\cup_{n\geq 1}{\mathcal{N}}_{n}. ∎

Lemma 11.3.

For all exceptional set 𝒩{\mathcal{N}} with respect to 𝕏{\mathbb{X}}, all n1n\geq 1 and all xE2nx\in E^{n}_{2}, we have xn(Xτnn𝒩)=1{\mathbb{P}}^{n}_{x}(X^{n}_{\tau_{n}}\notin{\mathcal{N}})=1.

Proof.

We fix 𝒩{\mathcal{N}} an exceptional set with respect to 𝕏{\mathbb{X}}, n1n\geq 1 and xE2nx\in E^{n}_{2}. For ε(0,1]\varepsilon\in(0,1], we write

xn(Xτnn𝒩)xn(τnε)+xn(τn>ε,Xτnn𝒩)=xn(τnε)+𝔼xn[1I{τn>ε}Xεnn(Xτnn𝒩)]{\mathbb{P}}^{n}_{x}(X^{n}_{\tau_{n}}\in{\mathcal{N}})\leq{\mathbb{P}}^{n}_{x}(\tau_{n}\leq\varepsilon)+{\mathbb{P}}^{n}_{x}(\tau_{n}>\varepsilon,X^{n}_{\tau_{n}}\in{\mathcal{N}})={\mathbb{P}}^{n}_{x}(\tau_{n}\leq\varepsilon)+{\mathbb{E}}^{n}_{x}[\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\tau_{n}>\varepsilon\}}{\mathbb{P}}^{n}_{X_{\varepsilon}^{n}}(X^{n}_{\tau_{n}}\in{\mathcal{N}})]

by the Markov property. But by Lemma 11.2, for all yE2n𝒩0y\in E^{n}_{2}\setminus{\mathcal{N}}_{0}, the law of (Xtτnn)t0(X^{n}_{t\land\tau_{n}})_{t\geq 0} under yn{\mathbb{P}}^{n}_{y} is equal to the law of (Xtσn)t0(X_{t\land\sigma_{n}})_{t\geq 0} under yX{\mathbb{P}}^{X}_{y}. Since 𝒩0𝒩{\mathcal{N}}_{0}\cup{\mathcal{N}} is exceptional for 𝕏{\mathbb{X}}, we can find 𝒩𝒩0𝒩{\mathcal{N}}^{\prime}\supset{\mathcal{N}}_{0}\cup{\mathcal{N}} properly exceptional for 𝕏{\mathbb{X}} (see Subsection B.1). Hence for all yE2n𝒩y\in E^{n}_{2}\setminus{\mathcal{N}}^{\prime},

yn(Xτnn𝒩)yn(Xτnn𝒩)=yX(Xσn𝒩)=0.{\mathbb{P}}^{n}_{y}(X^{n}_{\tau_{n}}\in{\mathcal{N}})\leq{\mathbb{P}}^{n}_{y}(X^{n}_{\tau_{n}}\in{\mathcal{N}}^{\prime})={\mathbb{P}}^{X}_{y}(X_{\sigma_{n}}\in{\mathcal{N}}^{\prime})=0.

Since xn(Xεn)1{\mathbb{P}}^{n}_{x}\circ(X^{n}_{\varepsilon})^{-1} has a density by Lemma 11.2, we conclude that xn(Xεn𝒩)=0{\mathbb{P}}^{n}_{x}(X^{n}_{\varepsilon}\in{\mathcal{N}}^{\prime})=0 and thus that xn{\mathbb{P}}^{n}_{x}-a.s., we have Xεnn(Xτnn𝒩)=0{\mathbb{P}}^{n}_{X_{\varepsilon}^{n}}(X^{n}_{\tau_{n}}\in{\mathcal{N}})=0. All in all, we have proved that xn(Xτnn𝒩)xn(τnε){\mathbb{P}}^{n}_{x}(X^{n}_{\tau_{n}}\in{\mathcal{N}})\leq{\mathbb{P}}^{n}_{x}(\tau_{n}\leq\varepsilon), and it suffices to let ε0\varepsilon\to 0, since xn(τn>0)=1{\mathbb{P}}^{n}_{x}(\tau_{n}>0)=1 by continuity and since xE2nx\in E_{2}^{n}. ∎

Using Lemmas 11.2 and 11.3, it is slightly technical but not difficult to build from 𝕏{\mathbb{X}} and the family (𝕏n)n1({\mathbb{X}}^{n})_{n\geq 1} a 𝒳{\mathcal{X}_{\triangle}}-valued diffusion 𝕏~=(Ω~X,~X,(X~t)t0,(~xX)x𝒳)\tilde{\mathbb{X}}=(\tilde{\Omega}^{X},\tilde{\mathcal{M}}^{X},(\tilde{X}_{t})_{t\geq 0},(\tilde{\mathbb{P}}^{X}_{x})_{x\in{\mathcal{X}}_{\triangle}}) such that \bullet for all x𝒳𝒩0x\in{\mathcal{X}_{\triangle}}\setminus{\mathcal{N}}_{0}, the law of (X~t)t0(\tilde{X}_{t})_{t\geq 0} under ~xX\tilde{\mathbb{P}}^{X}_{x} equals the law of (Xt)t0(X_{t})_{t\geq 0} under xX{\mathbb{P}}^{X}_{x}, \bullet for all x𝒩0x\in{\mathcal{N}}_{0}, setting n=1+max(1/dx,x)n=1+\lfloor\max(1/d_{x},||x||)\rfloor (so that xE2nx\in E^{n}_{2}), the law of (X~tσ~n)t0(\tilde{X}_{t\land\tilde{\sigma}_{n}})_{t\geq 0} under ~xX\tilde{\mathbb{P}}^{X}_{x} is the same as that of (Xtτnn)t0(X^{n}_{t\land\tau_{n}})_{t\geq 0} under xn{\mathbb{P}}^{n}_{x} and the law of (X~σ~n+t)t0(\tilde{X}_{\tilde{\sigma}_{n}+t})_{t\geq 0} under ~xX\tilde{\mathbb{P}}^{X}_{x} conditionally on ~σ~nX\tilde{\mathcal{M}}^{X}_{\tilde{\sigma}_{n}} equals the law of (Xt)t0(X_{t})_{t\geq 0} under X~σnX{\mathbb{P}}^{X}_{\tilde{X}_{\sigma_{n}}}. We have used the notation σ~n=inf{t>0:X~tE2n}\tilde{\sigma}_{n}=\inf\{t>0:\tilde{X}_{t}\notin E^{n}_{2}\} and ~tX=σ(X~s:s[0,t])\tilde{\mathcal{M}}^{X}_{t}=\sigma(\tilde{X}_{s}:s\in[0,t]).

Remark 11.4.

For all xE2x\in E_{2}, setting n=1+max(1/dx,x)n=1+\lfloor\max(1/d_{x},||x||)\rfloor, the law of (X~tσ~n)t0(\tilde{X}_{t\land\tilde{\sigma}_{n}})_{t\geq 0} under ~xX\tilde{\mathbb{P}}^{X}_{x} is the same as that of (Xtτnn)t0(X^{n}_{t\land\tau_{n}})_{t\geq 0} under xn{\mathbb{P}}^{n}_{x}.

Proof.

This follows from Lemma 11.2 when xE2𝒩0x\in E_{2}\setminus{\mathcal{N}}_{0} and from the definition of 𝕏~\tilde{\mathbb{X}} otherwise. ∎

We can finally give the

Proof of Proposition 1.2.

We fix N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta and we prove that 𝕏~\tilde{\mathbb{X}} defined above is a KS(θ,N)KS(\theta,N)-process. First, it is clear that 𝕏~\tilde{\mathbb{X}} is a QKS(θ,N)QKS(\theta,N)-process because 𝕏~\tilde{\mathbb{X}} is a 𝒳{\mathcal{X}_{\triangle}}-valued diffusion and since for all x𝒳𝒩0x\in{\mathcal{X}_{\triangle}}\setminus{\mathcal{N}}_{0}, the law of (X~t)t0(\tilde{X}_{t})_{t\geq 0} under ~xX\tilde{\mathbb{P}}^{X}_{x} equals the law of (Xt)t0(X_{t})_{t\geq 0} under xX{\mathbb{P}}^{X}_{x}, with 𝒩0{\mathcal{N}}_{0} exceptional for 𝕏{\mathbb{X}}. It remains to prove that for all xE2x\in E_{2}, all t>0t>0 and all Lebesgue-null A(2)NA\subset({\mathbb{R}}^{2})^{N}, we have ~xX(X~tA)=0\tilde{\mathbb{P}}^{X}_{x}(\tilde{X}_{t}\in A)=0. We set n=1+max(1/dx,x)n=1+\lfloor\max(1/d_{x},||x||)\rfloor and write, for any ε(0,t)\varepsilon\in(0,t),

~xX(X~tA)~xX(σ~n>ε,X~tA)+~xX(σ~nε)=𝔼~xX[1I{σ~n>ε}~X~εX(X~tεA)]+~xX(σ~nε).\tilde{\mathbb{P}}^{X}_{x}(\tilde{X}_{t}\in A)\leq\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}>\varepsilon,\tilde{X}_{t}\in A)+\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}\leq\varepsilon)=\tilde{\mathbb{E}}^{X}_{x}[\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\tilde{\sigma}_{n}>\varepsilon\}}\tilde{\mathbb{P}}^{X}_{\tilde{X}_{\varepsilon}}(\tilde{X}_{t-\varepsilon}\in A)]+\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}\leq\varepsilon).

Since 𝕏~\tilde{\mathbb{X}} is μ\mu-symmetric (because it is a QKS(θ,N)QKS(\theta,N)-process), since P~tε11\tilde{P}_{t-\varepsilon}1\leq 1, where P~t\tilde{P}_{t} is the semi-group of 𝕏~\tilde{\mathbb{X}} and since AA is Lebesgue-null,

(2)N~y(X~tεA)μ(dy)μ(A)=0.\int_{({\mathbb{R}}^{2})^{N}}\tilde{\mathbb{P}}_{y}(\tilde{X}_{t-\varepsilon}\in A)\mu({\rm d}y)\leq\mu(A)=0.

Hence there is a Lebesgue-null subset BB of (2)N({\mathbb{R}}^{2})^{N} (depending on tεt-\varepsilon) such that ~y(X~tεA)=0\tilde{\mathbb{P}}_{y}(\tilde{X}_{t-\varepsilon}\in A)=0 for every y(2)NBy\in({\mathbb{R}}^{2})^{N}\setminus B. We conclude that

~xX(X~tA)~xX(σ~n>ε,X~εB)+~xX(σ~nε)=xn(τn>ε,XεnB)+~xX(σ~nε),\tilde{\mathbb{P}}^{X}_{x}(\tilde{X}_{t}\in A)\leq\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}>\varepsilon,\tilde{X}_{\varepsilon}\in B)+\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}\leq\varepsilon)={\mathbb{P}}^{n}_{x}(\tau_{n}>\varepsilon,X_{\varepsilon}^{n}\in B)+\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}\leq\varepsilon),

where we finally used Remark 11.4. Since BB is Lebesgue-null, we deduce from Lemma 11.1 that xn(τn>ε,XεnB)=0{\mathbb{P}}^{n}_{x}(\tau_{n}>\varepsilon,X_{\varepsilon}^{n}\in B)=0. Thus ~xX(X~tA)~xX(σ~nε)\tilde{\mathbb{P}}^{X}_{x}(\tilde{X}_{t}\in A)\leq\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}\leq\varepsilon), which tends to 0 as ε0\varepsilon\to 0 because ~xX(σ~n>0)=1\tilde{\mathbb{P}}^{X}_{x}(\tilde{\sigma}_{n}>0)=1 by continuity. ∎

11.2. Final proofs

We fix θ>0\theta>0, N2N\geq 2 such that N>θN>\theta and a KS(θ,N)KS(\theta,N)-process 𝕏{\mathbb{X}}, which exists thanks to Subsection 11.1. We recall that E2nE^{n}_{2} was introduced in (2) and define, for all n1n\geq 1, σn=inf{t0:XtE2n}\sigma_{n}=\inf\{t\geq 0:X_{t}\notin E^{n}_{2}\}, as well as the σ\sigma-field

𝒢=n1σ(Xσn+t,t0).{\mathcal{G}}=\cap_{n\geq 1}\sigma(X_{\sigma_{n}+t},t\geq 0).
Lemma 11.5.

Fix A𝒢A\in{\mathcal{G}}. If xX(A)=0{\mathbb{P}}^{X}_{x}(A)=0 quasi-everywhere, then xX(A)=0{\mathbb{P}}^{X}_{x}(A)=0 for all xE2x\in E_{2}.

Proof.

We fix A𝒢A\in{\mathcal{G}} such that xX(A)=0{\mathbb{P}}^{X}_{x}(A)=0 quasi-everywhere. There is an exceptional set 𝒩{\mathcal{N}} such that for all xE2𝒩x\in E_{2}\setminus{\mathcal{N}}, xX(A)=0{\mathbb{P}}^{X}_{x}(A)=0. We now fix xE2x\in E_{2} and set n=1+max(1/dx,x)n=1+\lfloor\max(1/d_{x},||x||)\rfloor. For any ε(0,1]\varepsilon\in(0,1],

xX(A)xX(σnε)+xX[σn>ε,A].{\mathbb{P}}^{X}_{x}(A)\leq{\mathbb{P}}^{X}_{x}(\sigma_{n}\leq\varepsilon)+{\mathbb{P}}^{X}_{x}[\sigma_{n}>\varepsilon,A].

By the Markov property and since A𝒢σ(Xσn+t,t0)A\in{\mathcal{G}}\subset\sigma(X_{\sigma_{n}+t},t\geq 0), we get

xX[σn>ε,A]=𝔼xX[1I{σn>ε}XεX(A)].{\mathbb{P}}^{X}_{x}[\sigma_{n}>\varepsilon,A]={\mathbb{E}}^{X}_{x}[\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\sigma_{n}>\varepsilon\}}{\mathbb{P}}^{X}_{X_{\varepsilon}}(A)].

But the law of XεX_{\varepsilon} under xX{\mathbb{P}}^{X}_{x} has a density, so that xX(Xε𝒩)=0{\mathbb{P}}^{X}_{x}(X_{\varepsilon}\in{\mathcal{N}})=0, whence xX(XεX(A)=0)=1{\mathbb{P}}^{X}_{x}({\mathbb{P}}^{X}_{X_{\varepsilon}}(A)=0)=1. Hence xX[σn>ε,A]=0{\mathbb{P}}^{X}_{x}[\sigma_{n}>\varepsilon,A]=0 and we end with xX(A)xX(τnε){\mathbb{P}}^{X}_{x}(A)\leq{\mathbb{P}}^{X}_{x}(\tau_{n}\leq\varepsilon). As usual, we conclude that xX(A)=0{\mathbb{P}}^{X}_{x}(A)=0 by letting ε0\varepsilon\to 0. ∎

We are now ready to give the

Proof of Proposition 1.3.

Let θ(0,2)\theta\in(0,2) and N2N\geq 2. Since our KS(θ,N)KS(\theta,N)-process 𝕏{\mathbb{X}} is a QKS(θ,N)QKS(\theta,N)-process, we know from Proposition 8.1-(i) that xX(ζ=)=1{\mathbb{P}}^{X}_{x}(\zeta=\infty)=1 quasi-everywhere. We want to prove that xX(ζ=)=1{\mathbb{P}}^{X}_{x}(\zeta=\infty)=1 for all xE2x\in E_{2}. By Lemma 11.5, it thus suffices to check that {ζ=}\{\zeta=\infty\} belongs to 𝒢{\mathcal{G}}, which is not hard since for each n1n\geq 1,

{ζ=}={Xt𝒳 for all t0}={Xt𝒳 for all tσn}σ(Xσn+t,t0).\{\zeta=\infty\}=\{X_{t}\in{\mathcal{X}}\hbox{ for all }t\geq 0\}=\{X_{t}\in{\mathcal{X}}\hbox{ for all }t\geq\sigma_{n}\}\in\sigma(X_{\sigma_{n}+t},t\geq 0).

For the second equality, we used that XtE¯2n𝒳X_{t}\in\bar{E}^{n}_{2}\subset{\mathcal{X}} for all t[0,σn]t\in[0,\sigma_{n}] by definition. ∎

Proof of Theorem 1.5.

Let θ2\theta\geq 2 and N>3θN>3\theta. Since our KS(θ,N)KS(\theta,N)-process 𝕏{\mathbb{X}} is a QKS(θ,N)QKS(\theta,N)-process, we know from Section 10 that all the conclusions of Theorem 1.5 hold quasi-everywhere. In other words, xX(A)=1{\mathbb{P}}_{x}^{X}(A)=1 quasi-everywhere, where AA is the event on which we have ζ<\zeta<\infty, Xζ=limtζXt(2)NX_{\zeta-}=\lim_{t\to\zeta-}X_{t}\in({\mathbb{R}}^{2})^{N}, there is K0[[1,N]]K_{0}\in[\![1,N]\!] with cardinal |K0|=k0|K_{0}|=k_{0} such that there is a K0K_{0}-collision in the configuration XζX_{\zeta-}, etc. We want to prove that xX(A)=1{\mathbb{P}}^{X}_{x}(A)=1 for all xE2x\in E_{2}. By Lemma 11.5, it thus suffices to check that AA belongs to 𝒢{\mathcal{G}}. But for each n1n\geq 1, AA indeed belongs to σ(Xσn+t,t0)\sigma(X_{\sigma_{n}+t},t\geq 0), because no collision (nor explosion) may happen before getting out of E2nE^{n}_{2}. ∎

We end this section with the following remark (that we will not use anywhere).

Remark 11.6.

Fix θ0\theta\geq 0 and N2N\geq 2 such that N>θN>\theta. Consider a KS(θ,N)KS(\theta,N) process 𝕏{\mathbb{X}} and define σ=inf{t0:XtE2}\sigma=\inf\{t\geq 0:X_{t}\notin E_{2}\}. For all xE2x\in E_{2}, there is some (tX)t0({\mathcal{M}}^{X}_{t})_{t\geq 0}-Brownian motion ((Bti)t0)i[[1,N]]((B^{i}_{t})_{t\geq 0})_{i\in[\![1,N]\!]} (of dimension 2N2N) under xX{\mathbb{P}}^{X}_{x} such that for all t[0,σ)t\in[0,\sigma), all i[[1,N]]i\in[\![1,N]\!],

(3) Xti=xi+BtiθNji0tXsiXsjXsiXsj2ds.X_{t}^{i}=x^{i}+B^{i}_{t}-\frac{\theta}{N}\sum_{j\neq i}\int_{0}^{t}\frac{X^{i}_{s}-X^{j}_{s}}{||X^{i}_{s}-X^{j}_{s}||^{2}}{\rm d}s.
Proof.

It of course suffices to prove the result during [0,σn)[0,\sigma_{n}), where σn=inf{t0:XtE2n}\sigma_{n}=\inf\{t\geq 0:X_{t}\notin E_{2}^{n}\}. For any xE2nx\in E_{2}^{n} and for a given Brownian motion, the solutions to (3) and (1) classically coincide while they remain E2nE_{2}^{n}, because their drift coefficients coincide and are smooth inside E2nE_{2}^{n}. Hence, recalling the notation of Subsection 11.1, it suffices to prove that the semi-groups Pt(x,)P_{t}(x,\cdot) and Ptn(x,)P_{t}^{n}(x,\cdot) of the Markov processes 𝕏{\mathbb{X}} and 𝕏n{\mathbb{X}}^{n} killed when getting out of E2nE_{2}^{n} coincide for all xE2nx\in E_{2}^{n}. By Lemma 11.2, there is an exceptional set 𝒩0{\mathcal{N}}_{0} such that Pt(x,)=Ptn(x,)P_{t}(x,\cdot)=P_{t}^{n}(x,\cdot) for all xE2n𝒩0x\in E_{2}^{n}\setminus{\mathcal{N}}_{0}. We next fix xE2nx\in E_{2}^{n}. For any ε(0,t)\varepsilon\in(0,t), using that Pε(x,)P_{\varepsilon}(x,\cdot) has a density and that 𝒩0{\mathcal{N}}_{0} is Lebesgue-null, we easily deduce that Pt(x,)=(PεPtε)(x,)=(PεPtεn)(x,)P_{t}(x,\cdot)=(P_{\varepsilon}P_{t-\varepsilon})(x,\cdot)=(P_{\varepsilon}P^{n}_{t-\varepsilon})(x,\cdot). It is then not difficult, using that PtnP^{n}_{t} is Feller, to let ε0\varepsilon\to 0 and conclude that indeed, Pt(x,)=Ptn(x,)P_{t}(x,\cdot)=P^{n}_{t}(x,\cdot). ∎

Appendix A A few elementary computations

We recall that dθ,N(k)=(k1)(2θk/N)d_{\theta,N}(k)=(k-1)(2-\theta k/N) for k2k\geq 2 and give the

Proof of Lemma 1.1.

First, (3), which says that dθ,N(k)>0d_{\theta,N}(k)>0 if and only if k<k0=2N/θk<k_{0}=\lceil 2N/\theta\rceil, is clear. We next fix N>3θ6N>3\theta\geq 6, so that k0[[7,N]]k_{0}\in[\![7,N]\!] and dθ,N(2)=22θ/N(4/3,2)d_{\theta,N}(2)=2-2\theta/N\in(4/3,2). By concavity of x(x1)(2θx/N)x\to(x-1)(2-\theta x/N), it only remains to check that (i) dθ,N(3)2d_{\theta,N}(3)\geq 2, (ii) dθ,N(k03)2d_{\theta,N}(k_{0}-3)\geq 2, and (iii) dθ,N(k01)<2d_{\theta,N}(k_{0}-1)<2. We introduce a=2N/θ>6a=2N/\theta>6 and observe that dθ,N(k)=2a1(k1)(ak)d_{\theta,N}(k)=2a^{-1}(k-1)(a-k) and that k0=ak_{0}=\lceil a\rceil.

For (i), we write dθ,N(3)=4a1(a3)=412a1>2d_{\theta,N}(3)=4a^{-1}(a-3)=4-12a^{-1}>2 since a>6a>6.

For (ii), we have dθ,N(k03)=2a1(a4)(aa+3)d_{\theta,N}(k_{0}-3)=2a^{-1}(\lceil a\rceil-4)(a-\lceil a\rceil+3) and we need (a4)(aa+3)a(\lceil a\rceil-4)(a-\lceil a\rceil+3)\geq a. Writing a=n+αa=n+\alpha with an integer n6n\geq 6 and α(0,1]\alpha\in(0,1], we need that (n3)(2+α)n+α(n-3)(2+\alpha)\geq n+\alpha, and this holds true because 2(n3)n2(n-3)\geq n and (n3)αα(n-3)\alpha\geq\alpha.

For (iii), we write dθ,N(k01)=2a1(a2)(aa+1)2a1(a2)<2d_{\theta,N}(k_{0}-1)=2a^{-1}(\lceil a\rceil-2)(a-\lceil a\rceil+1)\leq 2a^{-1}(\lceil a\rceil-2)<2. ∎

We next study the reference measure of the Keller-Segel particle system.

Proposition A.1.

Let N2N\geq 2 and θ>0\theta>0 be such that N>θN>\theta. Recall that k0=2N/θk_{0}=\lceil 2N/\theta\rceil and the definition (4) of μ(dx)=𝐦(x)dx\mu({\rm d}x)={\mathbf{m}}(x){\rm d}x. (i) The measure μ\mu is Radon on Ek0E_{k_{0}}. (ii) If k0Nk_{0}\leq N, then μ\mu is not Radon on Ek0+1E_{k_{0}+1}.

Proof.

(i) To show that μ\mu is radon on Ek0E_{k_{0}}, we have to check that for all x=(x1,,xN)Ek0x=(x^{1},\dots,x^{N})\in E_{k_{0}}, which we now fix, there is an open set OxEk0O_{x}\subset E_{k_{0}} such that xOxx\in O_{x} and μ(Ox)<\mu(O_{x})<\infty. We choose Ox=i=1NB(xi,dx)O_{x}=\prod_{i=1}^{N}B(x^{i},d_{x}), where the balls are subsets of 2{\mathbb{R}}^{2} and where

dx=1min{xixj3:i,j[[1,N]] such that xixj}>0.d_{x}=1\land\min\Big{\{}\frac{\|x^{i}-x^{j}\|}{3}:i,j\in[\![1,N]\!]\mbox{ such that }x^{i}\neq x^{j}\Big{\}}>0.

We consider the partition K1,,KK_{1},\dots,K_{\ell} of [[1,N]][\![1,N]\!] such that for all pqp\neq q in [[1,]][\![1,\ell]\!], for all i,jKpi,j\in K_{p} and all kKqk\in K_{q}, xi=xjx^{i}=x^{j} and xixkx^{i}\neq x^{k}. Since xEk0x\in E_{k_{0}}, it holds that maxp[[1,]]|Kp|k01\max_{p\in[\![1,\ell]\!]}|K_{p}|\leq k_{0}-1. By definition of OxO_{x} and dxd_{x}, we see that for all yOxy\in O_{x}, for all pqp\neq q in [[1,]][\![1,\ell]\!], for all iKpi\in K_{p}, all jKqj\in K_{q},

yiyjxixjxiyixjyjxixj2dxdx.\|y^{i}-y^{j}\|\geq\|x^{i}-x^{j}\|-\|x^{i}-y^{i}\|-\|x^{j}-y^{j}\|\geq\|x^{i}-x^{j}\|-2d_{x}\geq d_{x}.

This implies that for some finite constant CC depending on xx, for all yOxy\in O_{x},

𝐦(y)=1ijNyiyjθ/NCp=1(i,jKp,ijyiyjθ/N).{\mathbf{m}}(y)=\prod_{1\leq i\neq j\leq N}||y^{i}-y^{j}||^{-\theta/N}\leq C\prod_{p=1}^{\ell}\Big{(}\prod_{i,j\in K_{p},i\neq j}||y^{i}-y^{j}||^{-\theta/N}\Big{)}.

Recall now that μ(dy)=𝐦(y)dy\mu({\rm d}y)={\mathbf{m}}(y){\rm d}y and that we want to show that μ(Ox)<\mu(O_{x})<\infty. Since xi=xjx^{i}=x^{j} for all i,jKpi,j\in K_{p} and all p[[1,]]p\in[\![1,\ell]\!], since |Kp|k01|K_{p}|\leq k_{0}-1, dx1d_{x}\leq 1 and by a translation argument, we are reduced to show that for any n[[2,k01]]n\in[\![2,k_{0}-1]\!], (when k0>Nk_{0}>N, one could study only n[[2,N]]n\in[\![2,N]\!])

In=(B(0,1))n(1ijnyiyjθ/N)dy1dyn<.I_{n}=\int_{(B(0,1))^{n}}\Big{(}\prod_{1\leq i\neq j\leq n}\|y^{i}-y^{j}\|^{-\theta/N}\Big{)}{\rm d}y^{1}\dots\rm dy^{n}<\infty.

We fix n[[2,k01]]n\in[\![2,k_{0}-1]\!] and show that In<I_{n}<\infty. Since u2|u1u2|\|u\|^{2}\geq|u_{1}u_{2}| for all u=(u1,u2)2u=(u_{1},u_{2})\in{\mathbb{R}}^{2}, we have InJn2I_{n}\leq J_{n}^{2}, where

Jn=[1,1]n(1ijn|titj|θ/(2N))dt1dtn.J_{n}=\int_{[-1,1]^{n}}\Big{(}\prod_{1\leq i\neq j\leq n}|t^{i}-t^{j}|^{-\theta/(2N)}\Big{)}{\rm d}t^{1}\dots\rm dt^{n}.

But for all t1,,tnt^{1},\dots,t^{n}\in{\mathbb{R}},

1ijn|titj|θ/(2N)=i=1n(j=1,jin|titj|θ/(2N))1ni=1nj=1,jin|titj|θn/(2N)\displaystyle\prod_{1\leq i\neq j\leq n}|t^{i}-t^{j}|^{-\theta/(2N)}=\prod_{i=1}^{n}\Big{(}\prod_{j=1,j\neq i}^{n}|t^{i}-t^{j}|^{-\theta/(2N)}\Big{)}\leq\frac{1}{n}\sum_{i=1}^{n}\prod_{j=1,j\neq i}^{n}|t^{i}-t^{j}|^{-\theta n/(2N)}

by the inequality of arithmetic and geometric means. Thus by symmetry,

Jn[1,1]n(j=2n|t1tj|θn/(2N))dt1dtn=11(11|t1t2|θn/(2N)dt2)n1dt1.J_{n}\leq\int_{[-1,1]^{n}}\Big{(}\prod_{j=2}^{n}|t^{1}-t^{j}|^{-\theta n/(2N)}\Big{)}{\rm d}t^{1}\dots\rm dt^{n}=\int_{-1}^{1}\Big{(}\int_{-1}^{1}|t^{1}-t^{2}|^{-\theta n/(2N)}{\rm d}t^{2}\Big{)}^{n-1}{\rm d}t^{1}.

Consequently,

Jn11(22|s|θn/(2N)ds)n1dt1.J_{n}\leq\int_{-1}^{1}\Big{(}\int_{-2}^{2}|s|^{-\theta n/(2N)}{\rm d}s\Big{)}^{n-1}{\rm d}t^{1}.

Since nk01=2N/θ1<2N/θn\leq k_{0}-1=\lceil 2N/\theta\rceil-1<2N/\theta, we have θn/(2N)<1\theta n/(2N)<1, so that Jn<J_{n}<\infty, whence In<I_{n}<\infty.

(ii) We next assume that k0[[2,N]]k_{0}\in[\![2,N]\!]. To prove that μ\mu is not radon on Ek0+1E_{k_{0}+1}, we show that μ(K)=\mu(K)=\infty for the compact subset

K=i=1k0B¯(0,1)×k=k0+1NB¯((2k,0),1/2)K=\prod_{i=1}^{k_{0}}\overline{B}(0,1)\times\prod_{k=k_{0}+1}^{N}\overline{B}((2k,0),1/2)

of Ek0+1E_{k_{0}+1}. All the balls in the previous formula are balls of 2{\mathbb{R}}^{2}. For x=(x1,,xN)Kx=(x^{1},\dots,x^{N})\in K, it holds that xk0+1,,xNx^{k_{0}+1},\dots,x^{N} are far from each other and far from x1,,xk0x^{1},\dots,x^{k_{0}}, which explains that KK is indeed compact in Ek0+1E_{k_{0}+1}. There is a positive constant c>0c>0 such that for all xKx\in K,

𝐦(x)=1ijNxixjθ/Nc1ijk0xixjθ/N,{\mathbf{m}}(x)=\prod_{1\leq i\neq j\leq N}||x^{i}-x^{j}||^{-\theta/N}\geq c\prod_{1\leq i\neq j\leq k_{0}}||x^{i}-x^{j}||^{-\theta/N},

whence, the value of c>0c>0 being allowed to vary,

μ(K)c(B(0,1))k0(1ijk0xixjθ/N)dx1dxk0.\mu(K)\geq c\int_{(B(0,1))^{k_{0}}}\Big{(}\prod_{1\leq i\neq j\leq k_{0}}\|x^{i}-x^{j}\|^{-\theta/N}\Big{)}{\rm d}x^{1}\dots\rm dx^{k_{0}}.

We now observe that

A={x=(x1,,xk0):x1,x2B(0,1/3),i{1,2},xiB(x1,x1x2)}(B(0,1))k0A=\{x=(x^{1},\dots,x^{k_{0}}):x^{1},x^{2}\in B(0,1/3),\;\forall i\notin\{1,2\},\;x^{i}\in B(x^{1},\|x^{1}-x^{2}\|)\}\subset(B(0,1))^{k_{0}}

and that for xAx\in A, we have xixjxix1+xjx12x1x2||x^{i}-x^{j}||\leq||x^{i}-x^{1}||+||x^{j}-x^{1}||\leq 2||x^{1}-x^{2}|| for all i,j=1,,k0i,j=1,\dots,k_{0}, from which

1ijk0xixjθ/Ncx1x2k0(k01)θ/N.\prod_{1\leq i\neq j\leq k_{0}}\|x^{i}-x^{j}\|^{-\theta/N}\geq c\|x^{1}-x^{2}\|^{-k_{0}(k_{0}-1)\theta/N}.

As a conclusion,

μ(K)\displaystyle\mu(K)\geq c(B(0,1/3))2x1x2k0(k01)θ/Ndx1dx2(B(x1,x1x2))k02dx3dxk0\displaystyle c\int_{(B(0,1/3))^{2}}\|x^{1}-x^{2}\|^{-k_{0}(k_{0}-1)\theta/N}{\rm d}x^{1}{\rm d}x^{2}\int_{(B(x_{1},\|x^{1}-x^{2}\|))^{k_{0}-2}}{\rm d}x^{3}\dots\rm dx^{k_{0}}
\displaystyle\geq c(B(0,1/3))2x1x2k0(k01)θ/N+2(k02)dx1dx2\displaystyle c\int_{(B(0,1/3))^{2}}\|x^{1}-x^{2}\|^{-k_{0}(k_{0}-1)\theta/N+2(k_{0}-2)}{\rm d}x^{1}{\rm d}x^{2}
\displaystyle\geq cB(0,1/3)uk0(k01)θ/N+2(k02)du,\displaystyle c\int_{B(0,1/3)}\|u\|^{-k_{0}(k_{0}-1)\theta/N+2(k_{0}-2)}{\rm d}u,

where we finally used the change of variables u=x1x2u=x^{1}-x^{2} and v=x1+x2v=x^{1}+x^{2}. This last integral diverges, because k0(k01)θ/N+2(k02)=dθ,N(k0)22-k_{0}(k_{0}-1)\theta/N+2(k_{0}-2)=d_{\theta,N}(k_{0})-2\leq-2, recall that dθ,N(k0)=(k01)(2k0θ/N)0d_{\theta,N}(k_{0})=(k_{0}-1)(2-k_{0}\theta/N)\leq 0 by definition of k0k_{0}. ∎

We need a similar result on the sphere 𝕊{\mathbb{S}} defined in Section 2, where γ:2(2)N\gamma:{\mathbb{R}}^{2}\to({\mathbb{R}}^{2})^{N} and Ψ:2×+×𝕊EN(2)N\Psi:{\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathbb{S}}\to E_{N}\subset({\mathbb{R}}^{2})^{N} were also introduced. First, we show an explicit link between μ(dx)=𝐦(x)dx\mu({\rm d}x)={\mathbf{m}}(x){\rm d}x and β(du)=𝐦(u)σ(du)\beta({\rm d}u)={\mathbf{m}}(u)\sigma({\rm d}u) defined in (4) and (1), that we use several times.

Lemma A.2.

We fix N2N\geq 2, θ>0\theta>0 and set ν=dθ,N(N)/21\nu=d_{\theta,N}(N)/2-1. For all Borel φ:(2)N+\varphi:({\mathbb{R}}^{2})^{N}\to{\mathbb{R}}_{+},

(2)Nφ(x)μ(dx)=122×+×𝕊φ(Ψ(z,r,u))rνdzdrβ(du).\int_{({\mathbb{R}}^{2})^{N}}\varphi(x)\mu({\rm d}x)=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}^{*}_{+}\times{\mathbb{S}}}\varphi(\Psi(z,r,u))r^{\nu}{\rm d}z{\rm d}r\beta({\rm d}u).
Proof.

Since H={y=(y1,,yN)(2)N:1Nyi=0}H=\{y=(y^{1},\dots,y^{N})\in({\mathbb{R}}^{2})^{N}:\sum_{1}^{N}y^{i}=0\} and since 𝐦{\mathbf{m}} is translation invariant,

(2)Nφ(x)μ(dx)=(2)Nφ(x)𝐦(x)dx=2×Hφ(γ(z)+y)𝐦(y)dzdy.\int_{({\mathbb{R}}^{2})^{N}}\varphi(x)\mu({\rm d}x)=\int_{({\mathbb{R}}^{2})^{N}}\varphi(x){\mathbf{m}}(x){\rm d}x=\int_{{\mathbb{R}}^{2}\times H}\varphi(\gamma(z)+y){\mathbf{m}}(y){\rm d}z{\rm d}y.

We next note that 𝕊{\mathbb{S}} is the (true) unit sphere of the (2N2)(2N-2)-dimensional Euclidean space HH and proceed to the substitution (,u)=(y,y/y)(\ell,u)=(\|y\|,y/||y||):

(2)Nφ(x)μ(dx)=2×+×𝕊φ(γ(z)+u)𝐦(u)2N3dzdσ(du).\int_{({\mathbb{R}}^{2})^{N}}\varphi(x)\mu({\rm d}x)=\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}\varphi(\gamma(z)+\ell u){\mathbf{m}}(\ell u)\ell^{2N-3}{\rm d}z{\rm d}\ell\sigma({\rm d}u).

We finally substitute =r\ell=\sqrt{r} and obtain

(2)Nφ(x)μ(dx)=122×+×𝕊φ(γ(z)+ru)𝐦(ru)rN2dzdrσ(du).\int_{({\mathbb{R}}^{2})^{N}}\varphi(x)\mu({\rm d}x)=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}\varphi(\gamma(z)+\sqrt{r}u){\mathbf{m}}(\sqrt{r}u)r^{N-2}{\rm d}z{\rm d}r\sigma({\rm d}u).

But 𝐦(ru)rN2=rN2θ(N1)/2𝐦(u){\mathbf{m}}(\sqrt{r}u)r^{N-2}=r^{N-2-\theta(N-1)/2}{\mathbf{m}}(u) by (4) and β(du)=𝐦(u)σ(du)\beta({\rm d}u)={\mathbf{m}}(u)\sigma({\rm d}u), whence

(2)Nφ(x)μ(dx)=122×+×𝕊φ(Ψ(z,r,u))rN2θ(N1)/2dzdrβ(du).\int_{({\mathbb{R}}^{2})^{N}}\varphi(x)\mu({\rm d}x)=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}\varphi(\Psi(z,r,u))r^{N-2-\theta(N-1)/2}{\rm d}z{\rm d}r\beta({\rm d}u).

Since finally ν=dθ,N(N)/21=N2θ(N1)/2\nu=d_{\theta,N}(N)/2-1=N-2-\theta(N-1)/2, the conclusion follows. ∎

We can now study the measure β\beta on 𝕊{\mathbb{S}}.

Proposition A.3.

Let N2N\geq 2 and θ>0\theta>0 such that N>θN>\theta. Recall that k0=2N/θk_{0}=\lceil 2N/\theta\rceil. (i) The measure β\beta is Radon on 𝕊Ek0{\mathbb{S}}\cap E_{k_{0}}. (ii) If k0Nk_{0}\geq N, then β(𝕊)<\beta({\mathbb{S}})<\infty .

Proof.

We start with (i). For ε(0,1]\varepsilon\in(0,1], we introduce

𝒦ε={x(2)N:K[[1,N]] such that |K|k0, we have RK(x)ε}andε=𝒦ε𝕊.{\mathcal{K}}_{\varepsilon}=\{x\in({\mathbb{R}}^{2})^{N}:\forall K\subset[\![1,N]\!]\mbox{ such that }|K|\geq k_{0},\mbox{ we have }R_{K}(x)\geq\varepsilon\}\quad\hbox{and}\quad{\mathcal{L}}_{\varepsilon}={\mathcal{K}}_{\varepsilon}\cap{\mathbb{S}}.

Since 𝒦εB¯(0,1){\mathcal{K}}_{\varepsilon}\cap\overline{B}(0,1) is compact in Ek0E_{k_{0}}, with here B(0,1)B(0,1) the unit ball of (2)N({\mathbb{R}}^{2})^{N}, we know from Proposition A.1-(i) that μ(𝒦εB(0,1))<\mu({\mathcal{K}}_{\varepsilon}\cap B(0,1))<\infty. Now by Lemma A.2,

μ(𝒦εB(0,1))=122×+×𝕊1I{γ(z)+ru𝒦εB(0,1)}rνdzdrβ(du).\mu({\mathcal{K}}_{\varepsilon}\cap B(0,1))=\frac{1}{2}\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}}}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\gamma(z)+\sqrt{r}u\in{\mathcal{K}}_{\varepsilon}\cap B(0,1)\}}r^{\nu}{\rm d}z{\rm d}r\beta({\rm d}u).

But for (z,r,u)2×+×𝕊(z,r,u)\in{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}\times{\mathbb{S}},

γ(z)+ru𝒦εB(0,1)\gamma(z)+\sqrt{r}u\in{\mathcal{K}}_{\varepsilon}\cap B(0,1)     if and only if     uε/ru\in{\mathcal{L}}_{\varepsilon/r}   and   Nz2+r<1N||z||^{2}+r<1.

Indeed, RK(γ(z)+ru)=rRK(u)R_{K}(\gamma(z)+\sqrt{r}u)=rR_{K}(u) for all K[[1,N]]K\subset[\![1,N]\!] and γ(z)+ru2=1Nz+rui2=Nz2+r||\gamma(z)+\sqrt{r}u||^{2}=\sum_{1}^{N}||z+\sqrt{r}u^{i}||^{2}=N||z||^{2}+r because 1Nui=0\sum_{1}^{N}u^{i}=0 and 1Nui2=1\sum_{1}^{N}||u^{i}||^{2}=1. Thus

μ(𝒦εB(0,1))=2×+1I{Nz2+r<1}rνβ(ε/r)dzdr.\mu({\mathcal{K}}_{\varepsilon}\cap B(0,1))=\int_{{\mathbb{R}}^{2}\times{\mathbb{R}}_{+}}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{N||z||^{2}+r<1\}}r^{\nu}\beta({\mathcal{L}}_{\varepsilon/r}){\rm d}z{\rm d}r.

All this implies that for all ε(0,1]\varepsilon\in(0,1], for almost all r(0,1)r\in(0,1), β(ε/r)<\beta({\mathcal{L}}_{\varepsilon/r})<\infty. Since εε\varepsilon\to{\mathcal{L}}_{\varepsilon} is monotone, we conclude that β(ε)<\beta({\mathcal{L}}_{\varepsilon})<\infty for all ε(0,1]\varepsilon\in(0,1]. Since finally ε(0,1]ε=𝕊Ek0\cup_{\varepsilon\in(0,1]}{\mathcal{L}}_{\varepsilon}={\mathbb{S}}\cap E_{k_{0}} and since ε{\mathcal{L}}_{\varepsilon} is compact in 𝕊Ek0{\mathbb{S}}\cap E_{k_{0}} for each ε(0,1]\varepsilon\in(0,1], we conclude as desired that β\beta is Radon on 𝕊Ek0{\mathbb{S}}\cap E_{k_{0}}.

We next prove (ii). It holds that 𝕊EN{\mathbb{S}}\subset E_{N}, because for u𝕊u\in{\mathbb{S}}, we have R[[1,N]](u)=1R_{[\![1,N]\!]}(u)=1. Hence if k0Nk_{0}\geq N, then 𝕊ENEk0{\mathbb{S}}\subset E_{N}\subset E_{k_{0}} , whence 𝕊=𝕊Ek0{\mathbb{S}}={\mathbb{S}}\cap E_{k_{0}} and thus β\beta is Radon on 𝕊{\mathbb{S}} by point (i). Since finally 𝕊{\mathbb{S}} is compact, we conclude that β(𝕊)<\beta({\mathbb{S}})<\infty. ∎

Appendix B Markov processes and Dirichlet spaces

In a first subsection, we recall some classical definitions and results about Hunt processes, diffusions and Dirichlet spaces found in Fukushima-Oshima-Takeda [11]. In a second subsection, we mention a few results about martingales, times-changes, concatenation, killing and Girsanov transformation of Hunt processes found in [11] and elsewhere.

B.1. Main definitions and properties

Let EE be a locally compact separable metrizable space endowed with a Radon measure α\alpha such that Supp α=E\alpha=E. We set E=E{}{E_{\triangle}}=E\cup\{\triangle\}, where \triangle is a cemetery point. See [11, Section A2] for the definition of a Hunt process 𝕐=(Ω,,(Yt)t0,(y)yE){\mathbb{Y}}=(\Omega,{\mathcal{M}},(Y_{t})_{t\geq 0},({\mathbb{P}}_{y})_{y\in{E_{\triangle}}}): it is a strong Markov process in its canonical filtration, y(Y0=y)=1{\mathbb{P}}_{y}(Y_{0}=y)=1 for all yEy\in{E_{\triangle}}, \triangle is an absorbing state, i.e. Yt=Y_{t}=\triangle for all t0t\geq 0 under {\mathbb{P}}_{\triangle}, and a few more technical properties are satisfied. The life-time of 𝕐{\mathbb{Y}} is defined by ζ=inf{t0:Yt=}\zeta=\inf\{t\geq 0:Y_{t}=\triangle\}.

Let us denote by Pt(y,dz)P_{t}(y,{\rm d}z) its transition kernel. Our Hunt process is said to be α\alpha-symmetric if EφPtψdα=EψPtφdα\int_{E}\varphi P_{t}\psi{\rm d}\alpha=\int_{E}\psi P_{t}\varphi{\rm d}\alpha for all measurable φ,ψ:E+\varphi,\psi:E\to{\mathbb{R}}_{+} and all t0t\geq 0, see [11, page 30]. The Dirichlet space (,)({\mathcal{E}},{\mathcal{F}}) of our Hunt process on L2(E,α)L^{2}(E,\alpha) is then defined, see [11, page 23], by

={φL2(E,α):limt01tEφ(Ptφφ)dα exists},\displaystyle{\mathcal{F}}=\Big{\{}\varphi\in L^{2}(E,\alpha):\lim_{t\to 0}\frac{1}{t}\int_{E}\varphi(P_{t}\varphi-\varphi){\rm d}\alpha\hbox{ exists}\Big{\}},
(φ,ψ)=limt01tEφ(Ptψψ)dαfor all φ,ψ.\displaystyle{\mathcal{E}}(\varphi,\psi)=-\lim_{t\to 0}\frac{1}{t}\int_{E}\varphi(P_{t}\psi-\psi){\rm d}\alpha\qquad\hbox{for all $\varphi,\psi\in{\mathcal{F}}$}.

The generator (𝒜,𝒟A)({\mathcal{A}},{\mathcal{D}}_{A}) of 𝕐{\mathbb{Y}} is defined as follows:

𝒟𝒜={φL2(E,α):limt01t(Ptφφ) exists in L2(E,α)},\displaystyle{\mathcal{D}}_{\mathcal{A}}=\Big{\{}\varphi\in L^{2}(E,\alpha):\lim_{t\to 0}\frac{1}{t}(P_{t}\varphi-\varphi)\hbox{ exists in $L^{2}(E,\alpha)$}\Big{\}},

and for φ𝒟𝒜\varphi\in{\mathcal{D}}_{\mathcal{A}}, we denote by 𝒜φL2(E,α){\mathcal{A}}\varphi\in L^{2}(E,\alpha) this limit. By [11, Pages 20-21], it holds that

(1) 𝒟𝒜={φ:hL2(E,α) such that ψ, we have (φ,ψ)=Ehψdα}{\mathcal{D}}_{{\mathcal{A}}}=\Big{\{}\varphi\in{\mathcal{F}}:\exists\;h\in L^{2}(E,\alpha)\mbox{ such that }\forall\;\psi\in{\mathcal{F}},\;\hbox{ we have }\;{\mathcal{E}}(\varphi,\psi)=-\int_{E}h\psi{\rm d}\alpha\Big{\}}

and in such a case 𝒜φ=h{\mathcal{A}}\varphi=h.

The one-point compactification E=E{}{E_{\triangle}}=E\cup\{\triangle\} of EE is endowed with the topology consisting of all the open sets of EE and of all the sets of the form Kc{}K^{c}\cup\{\triangle\} with KK compact in EE, see page [11, page 69]. Observe that for a E{E_{\triangle}}-valued sequence (xn)n0(x_{n})_{n\geq 0}, we have limnxn=x\lim_{n}x_{n}=x if and only if

\bullet either xEx\in E, xnEx_{n}\in E for all nn large enough, and limnxn=xE\lim_{n}x_{n}=x\in E in the usual sense; \bullet or x=x=\triangle and for all compact subset KK of EE, there is nKn_{K}\in{\mathbb{N}} such that for all nnKn\geq n_{K}, xnKx_{n}\notin K.

We say that our Hunt process is continuous if tYtt\to Y_{t} is continuous from +{\mathbb{R}}_{+} into E{E_{\triangle}}, where E{E_{\triangle}} is endowed with the one-point compactification topology. A continuous Hunt process is called a diffusion.

A Dirichlet space (,)({\mathcal{E}},{\mathcal{F}}) on L2(E,α)L^{2}(E,\alpha) is said to be regular if it has a core, see [11, page 6], i.e. a subset 𝒞Cc(E){\mathcal{C}}\subset C_{c}(E)\cap{\mathcal{F}} which is dense in {\mathcal{F}} for the norm φ=[Eφ2dα+(φ,φ)]1/2||\varphi||=[\int_{E}\varphi^{2}{\rm d}\alpha+{\mathcal{E}}(\varphi,\varphi)]^{1/2} and dense in Cc(E)C_{c}(E) for the uniform norm.

Observe two regular Dirichlet spaces (,)({\mathcal{E}},{\mathcal{F}}) and (,)({\mathcal{E}}^{\prime},{\mathcal{F}}^{\prime}) such that (φ,φ)=(φ,φ){\mathcal{E}}(\varphi,\varphi)={\mathcal{E}}^{\prime}(\varphi,\varphi) for all φ\varphi in a common core 𝒞{\mathcal{C}} are necessarily equal, i.e. ={\mathcal{F}}={\mathcal{F}}^{\prime} and ={\mathcal{E}}={\mathcal{E}}^{\prime}. This follows from the fact that by definition, see [11, page 5], a Dirichlet space is closed.

We say that a Borel set AA of EE is (Pt)t0(P_{t})_{t\geq 0}-invariant if for all φL2(E,α)\varphi\in L^{2}(E,\alpha), all t>0t>0 we have Pt(1IAφ)=1IAPtφP_{t}(\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{A}\varphi)=\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{A}P_{t}\varphi α\alpha-a.e, see [11, page 53]. According to [11, page 55], we say that (,)({\mathcal{E}},{\mathcal{F}}) is irreducible if for all (Pt)t0(P_{t})_{t\geq 0}-invariant set AA, we have either α(A)=0\alpha(A)=0 or α(EA)=0\alpha(E\setminus A)=0.

We say that (,)({\mathcal{E}},{\mathcal{F}}) is recurrent if for all nonnegative φL1(E,α)\varphi\in L^{1}(E,\alpha), for α\alpha-a.e. yEy\in E, we have 𝔼y[0φ(Ys)ds]{0,}{\mathbb{E}}_{y}[\int_{0}^{\infty}\varphi(Y_{s}){\rm d}s]\in\{0,\infty\}, see [11, page 55].

We finally say that (,)({\mathcal{E}},{\mathcal{F}}) is transient if for all nonnegative φL1(E,α)\varphi\in L^{1}(E,\alpha), for α\alpha-a.e. yEy\in E, we have 𝔼y[0φ(Ys)ds]<{\mathbb{E}}_{y}[\int_{0}^{\infty}\varphi(Y_{s}){\rm d}s]<\infty, with the convention that φ()=0\varphi(\triangle)=0, see [11, page 55].

By [11, Lemma 1.6.4 page 55], if (,)({\mathcal{E}},{\mathcal{F}}) is irreducible, then it is either recurrent or transient.

A Borel set 𝒩E{\mathcal{N}}\subset E is properly exceptional if α(𝒩)=0\alpha({\mathcal{N}})=0 and y(t0:Yt𝒩)=0{\mathbb{P}}_{y}(\exists t\geq 0:Y_{t}\in{\mathcal{N}})=0 for all yE𝒩y\in E\setminus{\mathcal{N}}, see [11, page 153]. A property is said to hold true quasi-everywhere if it holds true outside a properly exceptional set.

Remark B.1.

Two Hunt processes with the same Dirichlet space share the same quasi-everywhere notion, up to the restriction that the capacity of every compact set is finite, which is always the case in the present work.

Proof.

We fix a Hunt process 𝕐{\mathbb{Y}} and explain why its quasi-everywhere notion depends only on its Dirichlet space. A set 𝒩E{\mathcal{N}}\subset E is exceptional, see [11, page 152], if there exists a Borel set 𝒩~\tilde{\mathcal{N}} such that 𝒩𝒩~{\mathcal{N}}\subset\tilde{\mathcal{N}} and y(t0:Yt𝒩~)=0{\mathbb{P}}_{y}(\exists t\geq 0:Y_{t}\in\tilde{\mathcal{N}})=0 for α\alpha-a.e. yEy\in E. A properly exceptional set is clearly exceptional and [11, Theorem 4.1.1 page 155] tells us that any exceptional set is included in a properly exceptional set. Thus, a property is true quasi-everywhere if and only if it holds true outside an exceptional set. Next, [11, Theorem 4.2.1-(ii) page 161] tells us that a set 𝒩{\mathcal{N}} is exceptional if and only if its capacity is 0, where the capacity of 𝒩E{\mathcal{N}}\subset E is entirely defined from the Dirichlet space. And for [11, Theorem 4.2.1-(ii) page 161] to apply, one needs that the capacity of all compact sets is finite. ∎

B.2. Toolbox

We start with martingales.

Lemma B.2.

Let EE be a locally compact separable metrizable space endowed with a Radon measure α\alpha such that Supp α=E\alpha=E, and (Ω,,(Zt)t0,(z)zE)(\Omega,{\mathcal{M}},(Z_{t})_{t\geq 0},({\mathbb{P}}_{z})_{z\in{E_{\triangle}}}) a α\alpha-symmetric E{E_{\triangle}}-valued diffusion with regular Dirichlet space (,)\left(\mathcal{E},\mathcal{F}\right) on L2(E,α)L^{2}(E,\alpha) and generator (𝒜,𝒟𝒜)({\mathcal{A}},{\mathcal{D}}_{\mathcal{A}}). Assume that φ:E\varphi:E\mapsto{\mathbb{R}} belongs to 𝒟𝒜{\mathcal{D}}_{\mathcal{A}} and that both φ\varphi and 𝒜φ{\mathcal{A}}\varphi are bounded. Define

Mtφ=φ(Zt)φ(Z0)0t𝒜φ(Zs)ds,M^{\varphi}_{t}=\varphi(Z_{t})-\varphi(Z_{0})-\int_{0}^{t}{\mathcal{A}}\varphi(Z_{s}){\rm d}s,

with the convention that φ()=𝒜φ()=0\varphi(\triangle)={\mathcal{A}}\varphi(\triangle)=0. Quasi-everywhere, (Mtφ)t0(M^{\varphi}_{t})_{t\geq 0} is a z{\mathbb{P}}_{z}-martingale in the canonical filtration of (Zt)t0(Z_{t})_{t\geq 0}.

This can be found in [11, page 332]. There the assumption on φ\varphi is that there is ff bounded and measurable such that φ=R1f\varphi=R_{1}f, i.e. φ=(I𝒜)1f\varphi=(I-{\mathcal{A}})^{-1}f, which simply means that φ𝒜φ\varphi-{\mathcal{A}}\varphi is bounded. Also, the conclusion is that (Mtφ)t0(M^{\varphi}_{t})_{t\geq 0} is a MAF, which indeed implies that (Mtφ)t0(M^{\varphi}_{t})_{t\geq 0} is a martingale, see [11, page 243].

Next, we deal with time-changes.

Lemma B.3.

Let EE be a CC^{\infty}-manifold, α\alpha a Radon measure on EE such that Supp(α)=E{\rm Supp}(\alpha)=E, and (Ω,,(Zt)t0,(z)zE)(\Omega,{\mathcal{M}},(Z_{t})_{t\geq 0},({\mathbb{P}}_{z})_{z\in{E_{\triangle}}}) a α\alpha-symmetric E{E_{\triangle}}-valued diffusion with regular Dirichlet space (,)\left(\mathcal{E},\mathcal{F}\right) on L2(E,α)L^{2}(E,\alpha) with core Cc(E)C^{\infty}_{c}(E). We also fix g:E(0,)g:E\to(0,\infty) continuous and take the convention that g()=0g(\triangle)=0. We consider the time-change At=0tg(Zs)dsA_{t}=\int_{0}^{t}g(Z_{s}){\rm d}s and its generalized inverse ρt=inf{s>0:As>t}\rho_{t}=\inf\{s>0:A_{s}>t\}. We introduce Yt=Zρt1I{ρt<}+1I{ρt=}Y_{t}=Z_{\rho_{t}}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\rho_{t}<\infty\}}+\triangle\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{\rho_{t}=\infty\}}. Then (Ω,,(Yt)t0,(y)yE)(\Omega,{\mathcal{M}},(Y_{t})_{t\geq 0},({\mathbb{P}}_{y})_{y\in{E_{\triangle}}}) is a gαg\alpha-symmetric E{E_{\triangle}}-valued diffusion with regular Dirichlet space (,)\left(\mathcal{E},\mathcal{F^{\prime}}\right) on L2(E,gα)L^{2}(E,g\alpha) with core Cc(E)C^{\infty}_{c}(E), i.e. {\mathcal{F}}^{\prime} is the closure of Cc(E)C^{\infty}_{c}(E) with respect to the norm [Eφ2gdα+(φ,φ)]1/2[\int_{E}\varphi^{2}g{\rm d}\alpha+{\mathcal{E}}(\varphi,\varphi)]^{1/2}.

Remark B.4.

If we apply the preceding result to the simple case where EE is an open subset of d{\mathbb{R}}^{d} and where (φ,φ)=dφ2dα{\mathcal{E}}(\varphi,\varphi)=\int_{{\mathbb{R}}^{d}}\|\nabla\varphi\|^{2}{\rm d}\alpha for all φCc(E)\varphi\in C^{\infty}_{c}(E), then when {\mathcal{E}} is seen as the Dirichlet form of a gαg\alpha-symmetric process, it may be better understood as (φ,φ)=dg1/2φ2gdα{\mathcal{E}}(\varphi,\varphi)=\int_{{\mathbb{R}}^{d}}\|g^{-1/2}\nabla\varphi\|^{2}g{\rm d}\alpha.

This lemma is nothing but a particular case of [11, Theorem 6.2.1 page 316], see also the few pages before. We only have to check that the Revuz measure in our case is gαg\alpha, i.e., see [11, (5.1.13) page 229], that for all bounded nonnegative measurable functions φ,ψ\varphi,\psi on EE, for all t>0t>0,

E𝔼x[0tφ(Zs)g(Zs)ds]ψ(x)α(dx)=0tE(PsZψ)φgdα,\int_{E}{\mathbb{E}}_{x}\Big{[}\int_{0}^{t}\varphi(Z_{s})g(Z_{s}){\rm d}s\Big{]}\psi(x)\alpha({\rm d}x)=\int_{0}^{t}\int_{E}(P^{Z}_{s}\psi)\varphi g{\rm d}\alpha,

where PtZP_{t}^{Z} is the semi-group of ZZ. The left hand side equals 0tEPsZ(φg)ψdα\int_{0}^{t}\int_{E}P_{s}^{Z}(\varphi g)\psi{\rm d}\alpha, so that the claim is obvious since ZZ is α\alpha-symmetric.

The following concatenation result can be found in Li-Ying [17, Proposition 3.2].

Lemma B.5.

Let EV,EWE_{V},E_{W} be two CC^{\infty}-manifolds, αV,αW\alpha_{V},\alpha_{W} be some Radon measures on EVE_{V} and EWE_{W} such that Supp(αV)=EV{\rm Supp}(\alpha_{V})=E_{V} and Supp(αW)=EW{\rm Supp}(\alpha_{W})=E_{W}. Let (ΩV,V,(Vt)t0,(vV)vEV{})(\Omega^{V},{\mathcal{M}}^{V},(V_{t})_{t\geq 0},({\mathbb{P}}_{v}^{V})_{v\in E_{V}\cup\{\triangle\}}) be a αV\alpha_{V}-symmetric (EV{})(E_{V}\cup\{\triangle\})-valued diffusion with regular Dirichlet space (V,V)\left(\mathcal{E}^{V},\mathcal{F}^{V}\right) on L2(EV,αV)L^{2}(E_{V},\alpha_{V}) with core Cc(EV)C^{\infty}_{c}(E_{V}). Consider (ΩW,W,(Wt)t0,(wW)wEW{})(\Omega^{W},{\mathcal{M}}^{W},(W_{t})_{t\geq 0},({\mathbb{P}}_{w}^{W})_{w\in E_{W}\cup\{\triangle\}}), a αW\alpha_{W}-symmetric (EW{})(E_{W}\cup\{\triangle\})-valued diffusion with regular Dirichlet space (W,W)\left(\mathcal{E}^{W},\mathcal{F}^{W}\right) on L2(EW,αW)L^{2}(E_{W},\alpha_{W}) with core Cc(EW)C^{\infty}_{c}(E_{W}). Introduce the measure α=αVαW\alpha=\alpha_{V}\otimes\alpha_{W} on E=EV×EWE=E_{V}\times E_{W}. We take the convention that (v,)=(,w)=(,)=(v,\triangle)=(\triangle,w)=(\triangle,\triangle)=\triangle for all vEVv\in E_{V}, all wEWw\in E_{W}. Moreover, we set (V,W)=σ({(Vt,Wt):t0}){\mathcal{M}}^{(V,W)}=\sigma(\{(V_{t},W_{t}):t\geq 0\}) and we define (v,w)(V,W)=vVwW{\mathbb{P}}^{(V,W)}_{(v,w)}={\mathbb{P}}_{v}^{V}\otimes{\mathbb{P}}_{w}^{W} if (v,w)EV×EW(v,w)\in E_{V}\times E_{W} and (V,W)=VW{\mathbb{P}}^{(V,W)}_{\triangle}={\mathbb{P}}_{\triangle}^{V}\otimes{\mathbb{P}}_{\triangle}^{W}. The process

(ΩV×ΩW,(V,W),(Vt,Wt)t0,((v,w)(V,W))(v,w)(EV×EW){})\Big{(}\Omega^{V}\times\Omega^{W},{\mathcal{M}}^{(V,W)},(V_{t},W_{t})_{t\geq 0},({\mathbb{P}}_{(v,w)}^{(V,W)})_{(v,w)\in(E_{V}\times E_{W})\cup\{\triangle\}}\Big{)}

is a E{E_{\triangle}}-valued α\alpha-symmetric diffusion , with regular Dirichlet space (,)({\mathcal{E}},{\mathcal{F}}) on L2(E,α)L^{2}(E,\alpha) with core Cc(E)C^{\infty}_{c}(E) and, for φCc(E)\varphi\in C^{\infty}_{c}(E),

(φ,φ)=EVW(φ(v,),φ(v,))αV(dv)+EWV(φ(,w),φ(,w))αW(dw).{\mathcal{E}}(\varphi,\varphi)=\int_{E_{V}}{\mathcal{E}}^{W}(\varphi(v,\cdot),\varphi(v,\cdot))\alpha_{V}({\rm d}v)+\int_{E_{W}}{\mathcal{E}}^{V}(\varphi(\cdot,w),\varphi(\cdot,w))\alpha_{W}({\rm d}w).

Observe that (V,W){\mathcal{M}}^{(V,W)} may be strictly smaller than VW{\mathcal{M}}^{V}\otimes{\mathcal{M}}^{W} due to the identification of all the cemetery points. Also, it actually holds true that VwW=vVW=VW{\mathbb{P}}_{\triangle}^{V}\otimes{\mathbb{P}}_{w}^{W}={\mathbb{P}}_{v}^{V}\otimes{\mathbb{P}}_{\triangle}^{W}={\mathbb{P}}_{\triangle}^{V}\otimes{\mathbb{P}}_{\triangle}^{W} on (V,W){\mathcal{M}}^{(V,W)} so that the choice (V,W)=VW{\mathbb{P}}^{(V,W)}_{\triangle}={\mathbb{P}}_{\triangle}^{V}\otimes{\mathbb{P}}_{\triangle}^{W} is arbitrary but legitimate.

The following killing result is a summary, adapted to our context, of Theorems 4.4.2 page 173 and 4.4.3-(i) page 174 in [11, Section 4.4].

Lemma B.6.

Let EE be a CC^{\infty}-manifold, let α\alpha be a Radon measure on EE such that Supp(α)=E{\rm Supp}(\alpha)=E, and let (Ω,,(Zt)t0,(z)zE)(\Omega,{\mathcal{M}},(Z_{t})_{t\geq 0},({\mathbb{P}}_{z})_{z\in{E_{\triangle}}}) be a α\alpha-symmetric E{E_{\triangle}}-valued diffusion with regular Dirichlet space (,)\left(\mathcal{E},\mathcal{F}\right) on L2(E,α)L^{2}(E,\alpha) with core Cc(E)C^{\infty}_{c}(E). Let OO be an open subset of EE and consider τO=inf{t0:XtO}\tau_{O}=\inf\{t\geq 0:X_{t}\notin O\}, with the convention that inf=\inf\emptyset=\infty. Then, setting

ZtO=Zt1I{t<τO}+1I{tτO},Z^{O}_{t}=Z_{t}\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{t<\tau_{O}\}}+\triangle\hbox{\rm 1}{\hskip-2.8pt}\hbox{\rm I}_{\{t\geq\tau_{O}\}},

(Ω,,(ZtO)t0,(z)zO{})(\Omega,{\mathcal{M}},(Z^{O}_{t})_{t\geq 0},({\mathbb{P}}_{z})_{z\in O\cup\{\triangle\}}) is a α|O\alpha|_{O}-symmetric O{}O\cup\{\triangle\}-valued diffusion with regular Dirichlet space (O,O)({\mathcal{E}}_{O},{\mathcal{F}}_{O}) on L2(O,α|O)L^{2}(O,\alpha|_{O}) with core Cc(O)C_{c}^{\infty}(O) and for φO\varphi\in{\mathcal{F}}_{O},

O(φ,φ)=(φ,φ).{\mathcal{E}}_{O}(\varphi,\varphi)={\mathcal{E}}(\varphi,\varphi).

Note that since OO is an open subset of the manifold EE and since the Hunt process is continuous, the regularity condition (4.4.6) of [11, Theorem 4.4.2 page 173] is obviously satisfied.

We finally give an adaptation of the Girsanov theorem in the context of Dirichlet spaces, which is a particular case of Chen-Zhang [5, Theorem 3.4].

Lemma B.7.

Let EE be an open subset of d{\mathbb{R}}^{d}, with d1d\geq 1, α\alpha be a Radon measure on EE such that Supp(α)=E{\rm Supp}(\alpha)=E and (Ω,,(Zt)t0,(z)zE)(\Omega,{\mathcal{M}},(Z_{t})_{t\geq 0},({\mathbb{P}}_{z})_{z\in{E_{\triangle}}}) be a α\alpha-symmetric E{E_{\triangle}}-valued diffusion with regular Dirichlet space (,)\left(\mathcal{E},\mathcal{F}\right) on L2(E,α)L^{2}(E,\alpha) with core Cc(E)C^{\infty}_{c}(E) such that for all φCc(E)\varphi\in C^{\infty}_{c}(E),

(φ,φ)=Eφ2dα.{\mathcal{E}}(\varphi,\varphi)=\int_{E}\|\nabla\varphi\|^{2}{\rm d}\alpha.

Let (𝒜,𝒟𝒜)({\mathcal{A}},{\mathcal{D}}_{{\mathcal{A}}}) stand for its generator. Let uu\in{\mathcal{F}} be bounded, such that for ϱ=eu\varrho=e^{u}, we have ϱ1𝒟𝒜\varrho-1\in{\mathcal{D}}_{\mathcal{A}} with 𝒜[ϱ1]{\mathcal{A}}[\varrho-1] is bounded. Set

Lϱt=ϱ(Zt)ϱ(Z0)exp(0t𝒜[ϱ1](Zs)ϱ(Zs)ds),L^{\varrho}_{t}=\frac{\varrho(Z_{t})}{\varrho(Z_{0})}\exp\Big{(}-\int_{0}^{t}\frac{{\mathcal{A}}\color[rgb]{0,0,0}[\varrho-1]\color[rgb]{0,0,0}(Z_{s})}{\varrho(Z_{s})}{\rm d}s\Big{)},

with the conventions that ϱ()=1\varrho(\triangle)=1 and 𝒜[ϱ1]()=0{\mathcal{A}}\color[rgb]{0,0,0}[\varrho-1]\color[rgb]{0,0,0}(\triangle)=0. Assume that ϱ\varrho is continuous on E{E_{\triangle}}. Then quasi-everywhere, (Ltϱ)t0(L_{t}^{\varrho})_{t\geq 0} is a bounded (t)t0({\mathcal{M}}_{t})_{t\geq 0}-martingale under z{\mathbb{P}}_{z}, where we have set t=σ({Zs:s[0,t]}){\mathcal{M}}_{t}=\sigma(\{Z_{s}:s\in[0,t]\}), and there exists a probability measure ~z\tilde{{\mathbb{P}}}_{z} on (Ω,)(\Omega,{\mathcal{M}}), such that for all t>0t>0, ~z=Ltϱz\tilde{{\mathbb{P}}}_{z}=L_{t}^{\varrho}\cdot{\mathbb{P}}_{z} on t{\mathcal{M}}_{t}.

Moreover (Ω,,(Zt)t0,(~z)zE)(\Omega,\!{\mathcal{M}},\!(Z_{t})_{t\geq 0},\!(\tilde{{\mathbb{P}}}_{z})_{z\in{E_{\triangle}}}) is a ϱ2α\varrho^{2}\!\alpha-symmetric E{E_{\triangle}}-valued diffusion with regular Dirichlet space (~,)(\tilde{{\mathcal{E}}},{\mathcal{F}}) on L2(E,ϱ2α)L^{2}(E,\varrho^{2}\alpha) such that for all φ\varphi\in{\mathcal{F}},

~(φ,φ)=12Eφ2ϱ2dα.\tilde{{\mathcal{E}}}(\varphi,\varphi)=\frac{1}{2}\int_{E}\|\nabla\varphi\|^{2}\varrho^{2}{\rm d}\alpha.

Actually, they speak of right processes in [5], but this is not an issue since we only consider continuous Hunt processes. Also, they assume that LϱL^{\varrho} is bounded from above and from below by some deterministic constants, on each compact time interval, but this is obvious under our assumptions on uu and 𝒜ϱ{\mathcal{A}}\varrho. Finally, their expression of LϱL^{\varrho} is different, see [5, pages 485-486]: first, they define MϱtM^{\varrho}_{t} as the martingale part of ϱ(Xt)\varrho(X_{t}). By Lemma B.2 (applied to ϱ1\varrho-1), we see that

Mϱt=ϱ(Zt)ϱ(Z0)0t𝒜[ϱ1](Zs)ds.M^{\varrho}_{t}=\varrho(Z_{t})-\varrho(Z_{0})-\int_{0}^{t}{\mathcal{A}}\color[rgb]{0,0,0}[\varrho-1]\color[rgb]{0,0,0}(Z_{s}){\rm d}s.

Then they put Mt=0t[ϱ(Zs)]1dMϱsM_{t}=\int_{0}^{t}[\varrho(Z_{s})]^{-1}{\rm d}M^{\varrho}_{s} and define LϱL^{\varrho} as

Lϱt=exp(Mt12Mt).L^{\varrho}_{t}=\exp\Big{(}M_{t}-\frac{1}{2}\langle M\rangle_{t}\Big{)}.

But by Itô’s formula, logϱ(Zt)=logϱ(Z0)+0t[ϱ(Zs)]1dMϱs+0t[ϱ(Zs)]1𝒜[ϱ1](Zs)ds120t[ϱ(Zs)]2dMϱs\log\varrho(Z_{t})=\log\varrho(Z_{0})+\int_{0}^{t}[\varrho(Z_{s})]^{-1}{\rm d}M^{\varrho}_{s}+\int_{0}^{t}[\varrho(Z_{s})]^{-1}{\mathcal{A}}\color[rgb]{0,0,0}[\varrho-1]\color[rgb]{0,0,0}(Z_{s}){\rm d}s-\frac{1}{2}\int_{0}^{t}[\varrho(Z_{s})]^{-2}{\rm d}\langle M^{\varrho}\rangle_{s}, whence logϱ(Zt)=logϱ(Z0)+Mt+0t[ϱ(Zs)]1𝒜[ϱ1](Zs)ds12Mt\log\varrho(Z_{t})=\log\varrho(Z_{0})+M_{t}+\int_{0}^{t}[\varrho(Z_{s})]^{-1}{\mathcal{A}}\color[rgb]{0,0,0}[\varrho-1]\color[rgb]{0,0,0}(Z_{s}){\rm d}s-\frac{1}{2}\langle M\rangle_{t}, so that Lϱt=exp(Mt12Mt)=[ϱ(Z0)]1ϱ(Zt)exp(0tϱ(Zs)1𝒜[ϱ1](Zs)ds)L^{\varrho}_{t}=\exp(M_{t}-\frac{1}{2}\langle M\rangle_{t})=[\varrho(Z_{0})]^{-1}\varrho(Z_{t})\exp(-\int_{0}^{t}\varrho(Z_{s})^{-1}{\mathcal{A}}\color[rgb]{0,0,0}[\varrho-1]\color[rgb]{0,0,0}(Z_{s}){\rm d}s) as desired.

References

  • [1] G. Arumugam, J. Tyagi, Keller-Segel Chemotaxis Models: A Review, Acta Appl. Math. 171 (2021), 6.
  • [2] A. Blanchet, J. Dolbeault, B. Perthame, Two-dimensional Keller-Segel model: optimal critical mass and qualitative properties of the solutions, Electron. J. Differential Equations 44 (2006), 32 pp.
  • [3] D. Bresch, P.E. Jabin, Z. Wang On mean-field limits and quantitative estimates with a large class of singular kernels: application to the Patlak-Keller-Segel model, C. R. Math. Acad. Sci. Paris 357 (2019), 708–720.
  • [4] P. Cattiaux, L. Pédèches, The 2-D stochastic Keller-Segel particle model: existence and uniqueness, ALEA, Lat. Am. J. Probab. Math. Stat. 13 (2016), 447–463.
  • [5] Z. Chen, T.S. Zhang, Girsanov and Feynman-Kac type transformations for symmetric Markov processes. Ann. Inst. H. Poincaré Probab. Statist. 38 (2002), 475–505.
  • [6] J. Dolbeault, C. Schmeiser, The two-dimensional Keller-Segel model after blow-up, Discrete Contin. Dyn. Syst. 25 (2009), 109–121.
  • [7] I. Fatkullin, A study of blow-ups in the Keller-Segel model of chemotaxis, Nonlinearity 26 (2013), 81–94.
  • [8] N. Fournier, B. Jourdain, Stochastic particle approximation of the Keller-Segel equation and two-dimensional generalization of Bessel processes, Ann. Appl. Probab. 27 (2017), 2807–2861.
  • [9] N. Fournier, M. Hauray, S. Mischler, Propagation of chaos for the 2D viscous vortex model, J. Eur. Math. Soc. 16 (2014), 1423–1466.
  • [10] M. Fukushima, From one dimensional diffusions to symmetric Markov processes, Stochastic Process. Appl. 120 (2010), 590–604.
  • [11] M. Fukushima, Y. Oshima, M. Takeda, Dirichlet forms and symmetric Markov processes. Second revised and extended edition. Walter de Gruyter, 2011.
  • [12] J. Haškovec, C. Schmeiser, Stochastic particle approximation for measure valued solutions of the 2D Keller-Segel system, J. Stat. Phys. 135, 133–151.
  • [13] J. Haškovec, C. Schmeiser, Convergence of a stochastic particle approximation for measure solutions of the 2D Keller-Segel system, Comm. Partial Differential Equations 36(6) (2011), 940–960.
  • [14] J.F. Jabir, D. Talay, M. Tomasevic, Mean-field limit of a particle approximation of the one-dimensional parabolic-parabolic Keller-Segel model without smoothing, Electron. Commun. Probab. 23 (2018), Paper No. 84.
  • [15] I. Karatzas, S.E. Shreve, Brownian motion and stochastic calculus. Second edition. Graduate Texts in Mathematics, 113. Springer-Verlag, New York, 1991.
  • [16] E.F. Keller, L.A. Segel, Initiation of slime mold aggregation viewed as an instability, J. Theor. Biol. 26 (1970), 399–415.
  • [17] L. Li,, J. Ying, Regular subspaces of Dirichlet forms. Festschrift Masatoshi Fukushima, 397–420, Interdiscip. Math. Sci., 17, World Sci. Publ., Hackensack, NJ, 2015.
  • [18] C. Olivera, A. Richard, M. Tomasevic, Particle approximation of the 2-d parabolic-elliptic Keller-Segel system in the subcritical regime, arXiv:2011.00537.
  • [19] H. Osada, Propagation of chaos for the two-dimensional Navier-Stokes equation, Proc. Japan Acad. Ser. A Math. Sci. 62 (1986), 8–11.
  • [20] C.S. Patlak, Random walk with persistence and external bias, Bull. Math. Biophys. 15 (1953), 311–338.
  • [21] D. Revuz, M. Yor, Continuous martingales and Brownian motion. Third Edition. Springer, 2005.
  • [22] A. Stevens, The derivation of chemotaxis equations as limit dynamics of moderately interacting stochastic many-particle systems, SIAM J. Appl. Math. 61 (2000), 183–212.
  • [23] T. Suzuki Exclusion of boundary blowup for 2D chemotaxis system provided with Dirichlet boundary condition for the Poisson part, J. Math. Pures Appl. 100 (2013), 347–367.
  • [24] J.J.L. Velazquez, Point dynamics in a singular limit of the Keller-Segel model. I. Motion of the concentration regions, SIAM J. Appl. Math., 64 (2004), 1198–1223.
  • [25] J.J.L. Velazquez, Point dynamics in a singular limit of the Keller-Segel model. II. Formation of the concentration regions, SIAM J. Appl. Math., 64 (2004), 1224–1248.